a precision measurement of the weak mixing angle in møller ... · apv = −158 ppb±21 ppb (stat)...

193
A Precision Measurement of the Weak Mixing Angle in Møller Scattering at Low Q 2 Thesis by G. Mark Jones In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy California Institute of Technology Pasadena, California 2004 (Defended May 19, 2004)

Upload: others

Post on 26-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

A Precision Measurement of the Weak Mixing Angle inMøller Scattering at Low Q2

Thesis by

G. Mark Jones

In Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

California Institute of Technology

Pasadena, California

2004

(Defended May 19, 2004)

Page 2: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

ii

c© 2004

G. Mark Jones

All Rights Reserved

Page 3: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

iii

Acknowledgements

I am greatly indebted to many people that have provided support, encouragement, insight,

and guidance to me over the course of the work presented in this thesis. I would like first to

thank my advisor, Emlyn Hughes, for allowing me the opportunity to be a part of his group

at Caltech. I was afforded the chance to receive an excellent education in the classroom

and in the laboratory, and for that I am truly grateful.

The success of the challenging E158 experiment is due to the many diverse contributions

of the collaborators. To each of them I owe a great deal of thanks. In particular, I would

like to acknowledge Krishna Kumar, Paul Souder and my advisor for their able leadership

at SLAC. Their immense physical insight and experience were crucial to the experiment. I

am also grateful for the tireless efforts of Yury Kolomensky and Michael Woods. Without

their keen ability, the experiment could not have succeeded. Also, both Yury and Mike

were patient teachers, a trait greatly appreciated by the students on the experiment. I am

thankful that I got to know Raymond Arnold, who reminded me in the midst of all the

chaos of the experiment that I shouldn’t forget that the study of physics is an interesting

and worthwhile endeavor. Also, I would like to acknowledge Michael Olson ”MO” for being

an excellent run coordinator and teaching me the importance of the National Football

League and the Packers. I am thankful to Zenon Szalata for all the assistance he gave me,

especially on those owl shifts where I would inevitiably have to call him at his home for

Page 4: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

iv

advice. I would also like to acknowledge John Weisend who did a superb job with the liquid

hydrogen target, and who also taught me how to remain calm in the midst of warning bells

and flashing lights. I would like to thank Dieter Walz, whose unbounded knowledge about

the workings of the laboratory and his interesting tales made many a shift go by quickly.

The construction of the apparatus in End Station A within the time constraints of the

experiment proved to be a truly formidable undertaking. I am grateful for the diligent

work of Ken Baird that saw the experiment go from an idea on paper to existing in the

real world. I would also like to thank Michael Racine and Carl Hudspeth, as well as the

rest of the craftsmen who built the experiment, whose skill and expertise in all manners of

construction allowed the apparatus to function as designed.

I owe many thanks to my fellow students and the post-docs, who worked with me on

the data analysis, and also helped in taking all those shifts. In particular, I want to thank

Carlos Arroyo, Klejda Bega, Mark Cooke, Waled Emam, Brian Humensky, Lisa Kaufman,

Peter Mastromarino, Kent Paschke, Jaideep Singh, Baris Tonguc, Brock Tweedie, Imran

Younus, and Antonin Vacheret. We went through a lot together, and I feel privileged to

have worked with each of them.

I am truly grateful for having been able to work with Clive Field on the luminosity

monitor. I was always in awe of the combination of his sharp mind and his immense

technical ability. He taught me how to systematically approach problems, and he showed

me new ways of thinking about things. I would also like to acknowledge the contribution

of Gholamali Mazaheri, who worked diligently on the electronics side of the detector.

I was fortunate to also work with many interesting people during my time at Caltech.

David Pripstein introduced me to the work in the lab, and he also helped me adjust to life

in graduate school in general. His outgoing personality and his taste for fine wines made

Page 5: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

v

my first year at Caltech memorable and enjoyable. It is no exaggeration to say that my

interactions with Dave as an undergraduate led me to choose Caltech for graduate school.

For that, I am very thankful. Steffen Jensen and Jaideep Singh were always able to make

working in the lab more enjoyable with their sense of humor. I want to thank Tina Pavlin

for being a good listener and making sure I didn’t focus too much on lab work. Also, many

thanks to Georgia Frueh for skillfully managing all of the graduate students in our group

with a smile.

My experiences at the University of Kentucky as an undergraduate greatly influenced

my choice to pursue the study of physics. In particular, Wolfgang Korsch was a very

positive influence on me. His love of physics, as well as his patient guidance, gave me

great encouragement. The lessons he gave me in the classes he taught and at Jefferson lab

proved to be invaluable in graduate school. I would also like to mention Valdis Zeps and

Tim Gorringe, two excellent physics professors I had as an undergraduate. Their skillful

instruction prepared me for the rigors of graduate school and instilled in me a desire to

learn more. I would also like to thank Jian-ping Chen at Jefferson lab, who made my time

there both enjoyable and enlightening.

I owe special thanks to my friends and family. Peter Mastromarino, with his sense of

humor and genuine good nature, proved to be a true friend throughout graduate school. I

feel very fortunate that we ended up on the same experiment together. I am thankful to

Ted “Theodore” Corcovilos for his friendship, as well as helping me survive classwork at

Caltech and teaching me to “chill.” Sean Stave and Aaron Skaggs also provided me with

much needed encouragement, probably more than they know. And finally, I would like to

give thanks to my family, Gary, Linda, and Chris, to whom I literally owe everything. They

have always given me unconditional support and encouragement. I hope they feel they are a

Page 6: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

vi

part of the work in this thesis, because I know that I could not have accomplished anything

without them.

Page 7: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

vii

Abstract

The electroweak theory has been probed to a high level of precision at the mass scale of

the Z0 through the joint contributions of LEP at CERN and the SLC at SLAC. The E158

experiment at SLAC complements these results by measuring the weak mixing angle at a

Q2 of 0.026 (GeV/c)2, far below the weak scale.

The experiment utilizes a 48 GeV longitudinally polarized electron beam on unpolarized

atomic electrons in a target of liquid hydrogen to measure the parity-violating asymmetry

APV in Møller scattering.

The tree-level prediction for APV is proportional to 1− 4 sin2 θW . Since sin2 θW ≈ 0.25,

the effect of radiative corrections is enhanced, allowing the E158 experiment to probe for

physics effects beyond the Standard Model at the TeV scale.

This work presents the results from the first two physics runs of the experiment, covering

data collected in the year 2002. The parity-violating asymmetry APV was measured to be

APV = −158 ppb± 21 ppb (stat) ± 17 ppb (sys). (1)

The result represents the first demonstration of parity violation in Møller scattering. The

observed value of APV corresponds to a measurement of the weak mixing angle of

sin2 θeffW = 0.2380 ± 0.0016 (stat) ± 0.0013 (sys), (2)

Page 8: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

viii

which is in good agreement with the theoretical prediction of

sin2 θeffW = 0.2385 ± 0.0006 (theory). (3)

Page 9: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

ix

Contents

Acknowledgements iii

Abstract vii

1 Introduction 1

1.1 Early Study of the Weak Force . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Parity Violation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.2 Electroweak Unification . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.3 LEP and SLC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2 The Role of the E158 Experiment . . . . . . . . . . . . . . . . . . . . . . . . 5

1.3 E158 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4 Experiment Timeline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Theory 8

2.1 The E158 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.2 Previous Low Q2 Electroweak Measurements . . . . . . . . . . . . . . . . . 9

2.2.1 NuTeV Experiment Overview . . . . . . . . . . . . . . . . . . . . . . 10

2.2.2 Atomic Parity Violation Overview . . . . . . . . . . . . . . . . . . . 11

2.3 E158 Experiment at Tree Level . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.4 Radiative Corrections to APV . . . . . . . . . . . . . . . . . . . . . . . . . . 14

Page 10: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

x

2.4.1 γ − Z0 Mixing Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.4.2 Heavy Box Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.4.3 γ − Z0 Box Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.4.4 ρ Term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.4.5 Definition of sin2θW (Q2) . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 E158 Beamline and Beam Monitoring 20

3.1 Polarized Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.1.1 Laser Bench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

3.1.2 Combiner Bench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.1.3 Wall Bench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.1.4 Polarized Gun . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3.1.5 Helicity Sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3.1.6 Beam Asymmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.2 Beam Position Monitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2.1 BPM Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2.2 BPM Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.3 Charge Monitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.3.1 Toroid Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.3.2 Toroid Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.4 Wire Array . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.5 Skew Quadrupole Magnet . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.6 Liquid Hydrogen Target . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3.7 Spectrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

Page 11: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xi

3.7.1 Dipole Chicane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.7.2 Main Acceptance Collimator . . . . . . . . . . . . . . . . . . . . . . 44

3.7.3 Quadrupole Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.7.4 Insertable Acceptance Collimator . . . . . . . . . . . . . . . . . . . . 46

3.7.5 eP Collimator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.7.6 Synchrotron Collimation and Photon Masks . . . . . . . . . . . . . . 49

4 Detectors 52

4.1 E158 Calorimeter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.1.1 Calorimeter Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.1.2 Calorimeter Electronics . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.1.3 Calorimeter Resolution . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.1.4 Calorimeter Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . 59

4.2 Polarimeter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.2.1 Polarimetry Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.2.2 Polarimeter Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.2.3 Polarimeter Systematic Uncertainties . . . . . . . . . . . . . . . . . . 64

4.2.4 Polarimeter Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.3 Profile Monitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.3.1 Profile Monitor Design . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.3.2 Profile Scan Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.4 Pion Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.4.1 Pion Detector Design . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.4.2 Pion Detector Resolution . . . . . . . . . . . . . . . . . . . . . . . . 70

Page 12: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xii

4.4.3 Pion Detector Electromagnetic Background . . . . . . . . . . . . . . 71

4.4.4 Ratio of Nπ to Nee . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.4.5 Pion Detector Results . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.5 Synchrotron Light Monitor . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.5.1 Synchrotron Asymmetry . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.5.2 SLM Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.5.3 SLM Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.5.4 SLM Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5 Luminosity Monitor 80

5.1 Detector Signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

5.2 Synchrotron Radiation Background . . . . . . . . . . . . . . . . . . . . . . . 83

5.3 Detector Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

5.4 Lumi Signals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5.5 Gas System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

5.6 Asymmetry Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5.7 Resolution Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.8 Target Boiling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.9 Synchrotron Radiation Suppression . . . . . . . . . . . . . . . . . . . . . . . 97

5.10 Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.10.1 Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.10.2 Measured Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

5.11 Missing Pulse Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.12 Lumi as a BPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

Page 13: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xiii

6 Asymmetry Analysis 107

6.1 Initial Processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6.2 Detector Channel Weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6.3 Regression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6.4 Beam Dithering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

6.5 Data Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

6.6 Møller Detector Asymmetry Analysis . . . . . . . . . . . . . . . . . . . . . . 116

6.6.1 Systematic Reversals . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.6.2 Beam Corrections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

6.7 Asymmetry Corrections and Uncertainties . . . . . . . . . . . . . . . . . . . 122

6.7.1 The Electron-Proton Scattering Correction . . . . . . . . . . . . . . 122

6.7.2 Beam Asymmetry Correction Systematic Uncertainties . . . . . . . . 126

6.7.2.1 First-Order Beam Correction Systematic Uncertainties . . 126

6.7.2.2 Higher-Order Beam Correction Systematic Uncertainties . 129

6.7.3 Dipole Asymmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

6.7.4 Spot Size Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.7.5 Pion Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.7.6 Synchrotron Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.7.7 Neutral Backgrounds . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

6.7.8 Scale Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

6.8 Determination of APV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

6.9 Luminosity Monitor Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

6.9.1 Lumi Beam Correction Systematic Uncertainty . . . . . . . . . . . . 143

6.9.2 Lumi Dipole Contamination . . . . . . . . . . . . . . . . . . . . . . . 147

Page 14: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xiv

6.9.3 Lumi Dilutions and Scale Factors . . . . . . . . . . . . . . . . . . . . 148

6.10 Lumi Asymmetry Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

7 The Weak Mixing Angle 150

7.1 Extraction of sin2θW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

7.2 Physics beyond the Standard Model . . . . . . . . . . . . . . . . . . . . . . 154

7.2.1 Electron Compositeness Limit . . . . . . . . . . . . . . . . . . . . . . 154

7.2.2 Scalar Doubly Charged Higgs Limit . . . . . . . . . . . . . . . . . . 156

7.2.3 Extra Z Boson Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

7.2.4 Oblique Parameter X Limit . . . . . . . . . . . . . . . . . . . . . . . 159

7.3 Future Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7.3.1 E158 Run III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7.3.2 QWeak . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.3.3 DIS-Parity at JLab . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.3.4 The Large Hadron Collider . . . . . . . . . . . . . . . . . . . . . . . 162

7.3.5 Next Linear Collider . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

7.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

Bibliography 164

Page 15: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xv

List of Figures

1.1 Fermi’s four-point interaction for β decay. . . . . . . . . . . . . . . . . . . . . 2

1.2 The W− incorporated into β decay. . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 The running of sin2θW with four-momentum transfer Q. . . . . . . . . . . . . 6

1.4 General diagram of the E158 apparatus. The regions of the E158 calorimeter

are named as follows: A.) In ring B.) Mid ring C.) Out ring and D.) eP detector. 6

2.1 The running of sin2θW with Q. . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2 Neutrino scattering diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3 Anti-neutrino scattering diagrams. . . . . . . . . . . . . . . . . . . . . . . . . 10

2.4 Tree-level contributions to the nuclear potential. . . . . . . . . . . . . . . . . 11

2.5 Tree-level contributions to APV . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.6 Photon-Z0 mixing diagrams and the W contribution to the anapole moment. 15

2.7 Heavy boson box diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.8 Photon-Z0 box diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.9 Anapole moment contribution from the Z0. . . . . . . . . . . . . . . . . . . . 17

3.1 Location of the polarized source room. . . . . . . . . . . . . . . . . . . . . . . 20

3.2 E158 optics configuration at the source. . . . . . . . . . . . . . . . . . . . . . 21

3.3 Helicity of the E158 electron beam. . . . . . . . . . . . . . . . . . . . . . . . 24

Page 16: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xvi

3.4 Integrated charge asymmetry measured near the source, spanning all of Run

1 and Run 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.5 Run 1 and Run 2 target region integrated charge asymmetry. . . . . . . . . . 26

3.6 Location of E158 beam monitors. . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.7 E158 beam position monitor. . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.8 BPM signals after mixing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.9 Angle and position BPM resolutions. . . . . . . . . . . . . . . . . . . . . . . . 30

3.10 Energy BPM resolutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.11 Toroid resolution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3.12 Toroid signal versus calibrator signal. . . . . . . . . . . . . . . . . . . . . . . 35

3.13 Wire array profiles for both X and Y axes. . . . . . . . . . . . . . . . . . . . 36

3.14 Detector spot size correlation coefficients. . . . . . . . . . . . . . . . . . . . . 37

3.15 Location of skew quadrupole magnet. . . . . . . . . . . . . . . . . . . . . . . 39

3.16 Effect of the skew quad on detector resolution. . . . . . . . . . . . . . . . . . 39

3.17 E158 cryotarget loop. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3.18 Cryotarget scattering chamber. . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.19 E158 spectrometer overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.20 Regions of the E158 calorimeter. . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.21 E158 dipole chicane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.22 First photon collimator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.23 Main acceptance collimator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.24 Effect of main acceptance collimator on signal flux. . . . . . . . . . . . . . . . 46

3.25 Profile scans with and without quadrupole magnets. . . . . . . . . . . . . . . 46

3.26 Insertable acceptance collimator. . . . . . . . . . . . . . . . . . . . . . . . . . 47

Page 17: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xvii

3.27 Profile scans with (left) and without (right) the insertable collimator. . . . . 48

3.28 Insertable eP collimator, top view. . . . . . . . . . . . . . . . . . . . . . . . . 49

3.29 The hatched area represents the coverage of the eP collimator on the face of

the E158 calorimeter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.30 Profile scans without (left) and with (right) the eP collimator in place. . . . 50

3.31 Drift pipe synchrotron and photon collimators. . . . . . . . . . . . . . . . . . 51

4.1 Overhead view of the detector locations in End Station A. . . . . . . . . . . 52

4.2 Overhead view of the movable detector cart. . . . . . . . . . . . . . . . . . . 53

4.3 Partially constructed E158 calorimeter. . . . . . . . . . . . . . . . . . . . . . 54

4.4 Calorimeter channel map. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.5 Light guide configuration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.6 E158 calorimeter lead shielding. . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.7 E158 calorimeter electronics diagram. . . . . . . . . . . . . . . . . . . . . . . 57

4.8 Møller detector asymmetry resolution. . . . . . . . . . . . . . . . . . . . . . . 58

4.9 Møller plus Out ring asymmetry resolution. . . . . . . . . . . . . . . . . . . . 58

4.10 eP ring asymmetry resolution. . . . . . . . . . . . . . . . . . . . . . . . . . . 59

4.11 Polarimeter asymmetry distribution for a single run. . . . . . . . . . . . . . . 62

4.12 Polarimeter position on the detector cart. . . . . . . . . . . . . . . . . . . . . 63

4.13 Quartz-tungsten sandwich of polarimeter. . . . . . . . . . . . . . . . . . . . . 64

4.14 Run II beam polarization measurements, spanning 1 month. . . . . . . . . . 66

4.15 Profile monitor schematic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.16 Quartz scanner schematic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.17 Radial scan produced with the profile monitor. . . . . . . . . . . . . . . . . . 68

4.18 Pion detector layout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

Page 18: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xviii

4.19 Pion detector asymmetry distribution, covering one run. . . . . . . . . . . . . 71

4.20 SLM location in the A-Line bend. . . . . . . . . . . . . . . . . . . . . . . . . 74

4.21 SLM layout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.22 Regressed SLM channel 2 asymmetry distribution, for one run. . . . . . . . . 77

4.23 SLM asymmetry results for Run I and Run II. . . . . . . . . . . . . . . . . . 78

5.1 Layout of End Station A for experiment E158. . . . . . . . . . . . . . . . . . 81

5.2 Components of lumi signal at face of detector. . . . . . . . . . . . . . . . . . 81

5.3 Contributions to lumi asymmetry. . . . . . . . . . . . . . . . . . . . . . . . . 82

5.4 GEANT simulation of photon flux at the lumi. . . . . . . . . . . . . . . . . . 84

5.5 A. Front view of one full lumi ring, with sensitivity between 7 and 10 cm. B.

Side view, depicting the two lumi rings and the aluminum showering material. 85

5.6 Individual chamber design, with signal plates shaded. . . . . . . . . . . . . . 85

5.7 Lumi electronics setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

5.8 Front lumi signal traces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

5.9 Back lumi signal traces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

5.10 Gas system configuration for lumi. . . . . . . . . . . . . . . . . . . . . . . . . 89

5.11 X and Y correlation coefficients. The averages are shown as straight lines. . . 90

5.12 dX and dY correlation coefficients The averages are shown as straight lines. . 91

5.13 Left: Raw lumi asymmetry distribution. Right: Regression-corrected lumi

asymmetry distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

5.14 Extracted boiling noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

5.15 Synchrotron radiation background levels by chamber. . . . . . . . . . . . . . 98

5.16 Lumi signal versus beam current. . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.17 Full range of results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

Page 19: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xix

5.18 Tight beam cuts subset of linearity results. . . . . . . . . . . . . . . . . . . . 103

5.19 Charge normalized lumi signal following a missing pulse. . . . . . . . . . . . 104

5.20 The lumi asymmetry distribution without (left) and with (right) a cut after a

missing pulse. RMS is the statistical width of the distribution, while Sigma

refers to the width determined from the fit. . . . . . . . . . . . . . . . . . . . 104

5.21 Calculated beam position using the lumi, versus position position at lumi

figured using angle and position bpms. . . . . . . . . . . . . . . . . . . . . . . 106

6.1 Møller detector asymmetry versus Y position asymmetry. . . . . . . . . . . . 109

6.2 Møller detector asymmetry versus Y position asymmetry after regression. . . 111

6.3 Møller detector resolution with and without regression. . . . . . . . . . . . . 111

6.4 Location of components used for beam dithering. . . . . . . . . . . . . . . . . 112

6.5 A lumi chamber responding to a position dither cycle. . . . . . . . . . . . . . 113

6.6 Run I Møller detector asymmetry versus slug. . . . . . . . . . . . . . . . . . . 117

6.7 Run II Møller detector asymmetry versus slug. . . . . . . . . . . . . . . . . . 118

6.8 Measured asymmetry for each energy-halfwave plate setting. . . . . . . . . . 119

6.9 Run I Møller detector asymmetry versus slug, sign flips suppressed. . . . . . 120

6.10 Run II Møller detector asymmetry versus slug, sign flips suppressed. . . . . . 121

6.11 Data and simulation comparison with normal running conditions. . . . . . . 123

6.12 Data and simulation comparison with the insertable acceptance collimator in

position. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

6.13 Profile scan with the Møller and electron-proton scattering simulation results

superimposed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

6.14 Asymmetry result from the eP detector in Run I. . . . . . . . . . . . . . . . . 126

6.15 Run I Out ring asymmetry data. . . . . . . . . . . . . . . . . . . . . . . . . . 130

Page 20: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xx

6.16 Run II Out ring asymmetry data. . . . . . . . . . . . . . . . . . . . . . . . . 130

6.17 Run I asymmetry plotted versus channel number for 46 GeV running. . . . . 135

6.18 Run I asymmetry plotted versus channel number for production data. . . . . 136

6.19 Run I lumi asymmetry data. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

6.20 Run II lumi asymmetry data. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

6.21 Run I and Run II lumi asymmetry with systematic reversal sign flips ignored. 144

6.22 Run I lumi asymmetry plotted versus channel number. . . . . . . . . . . . . . 147

7.1 Bremsstrahlung diagrams included in Fb(y). The crossed versions must also

be computed, for a total of 16 diagrams. . . . . . . . . . . . . . . . . . . . . . 151

7.2 Experimental results and the theoretical running of the weak mixing angle. . 153

7.3 The measured weak mixing angle evolved to the Z0 mass. . . . . . . . . . . . 154

7.4 Doubly charged Higgs particle exchange diagram. . . . . . . . . . . . . . . . 156

7.5 Muonium to anti-muonium conversion. . . . . . . . . . . . . . . . . . . . . . . 157

7.6 Present and future experiments used to map the running of the weak mixing

angle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

Page 21: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xxi

List of Tables

1.1 The E158 dataset boundaries. . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

3.1 Integrated asymmetries for Run 1 and Run 2. . . . . . . . . . . . . . . . . . . 26

3.2 BPM contribution to Møller detector resolution. . . . . . . . . . . . . . . . . 31

3.3 Run 1 Estimate of uncertainty on APV due to BPM corrections. . . . . . . . 32

3.4 Run 1 systematic uncertainty on APV due to spot size. . . . . . . . . . . . . 38

3.5 Run 2 systematic uncertainty on APV due to spot size. . . . . . . . . . . . . 38

4.1 Measured linearity of the In, Mid, and Out rings of the E158 calorimeter, for

Run I and Run II. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.2 Polarimetry uncertainties, relative to Pbeam. . . . . . . . . . . . . . . . . . . . 64

4.3 Pion detector asymmetry results. . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.4 Vertical beam polarization at the target. . . . . . . . . . . . . . . . . . . . . . 78

4.5 Synchrotron asymmetry correction for the Møller detector. . . . . . . . . . . 79

5.1 Contributions to the luminosity monitor asymmetry resolution. *Target boil-

ing is covered in Section 5.8. . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.2 List of target boiling data runs. . . . . . . . . . . . . . . . . . . . . . . . . . . 96

6.1 Regression coefficients for Run I. . . . . . . . . . . . . . . . . . . . . . . . . . 112

6.2 Beam corrections to ARaw of the Møller detector. . . . . . . . . . . . . . . . . 120

Page 22: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xxii

6.3 Comparison of regression and dithering results for the Møller detector asym-

metry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.4 Dilution factors f due to background eP scatters. R is the ratio of elastic to

inelastic eP signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

6.5 Diluted asymmetries fA due to the electron-proton scattering background.

All entries are in ppb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

6.6 Regression coefficients of three Mid detector weighting schemes. . . . . . . . 127

6.7 Sensitive pattern-weighted detectors for the regression beam parameters. . . 128

6.8 Run I first-order systematic uncertainties in the regression corrections to the

raw Møller detector asymmetry. . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.9 Run II first-order systematic uncertainties in the regression corrections to the

raw Møller detector asymmetry. . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.10 The higher-order systematic uncertainty computed for the Møller detector,

comparing the Møller detector (1) and the Out ring (2), assuming Equation 6.26.133

6.11 The higher-order systematic uncertainty computed for the Møller detector,

comparing the Møller detector (1) and the Out ring (2), using Equation 6.29. 134

6.12 The computed shifts in AMeasured due to the dipole amplitude from transverse

beam polarization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.13 Synchrotron correction calculated with the SLM. . . . . . . . . . . . . . . . . 138

6.14 The dilution factors f are dimensionless, while the asymmetry corrections fA

are given in units of ppb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

6.15 The dilution factors f are dimensionless, while the asymmetry corrections fA

are given in units of ppb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6.16 The parity-violating asymmetry in Møller scattering. . . . . . . . . . . . . . . 141

Page 23: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

xxiii

6.17 Comparison of the regression coefficients for two individual lumi channels and

the full detector. Channel 00 is at the top of the lumi ring, while channel 02

is on the right side. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

6.18 The composite monitor Out-Mid has enhanced energy sensitivity. . . . . . . . 145

6.19 Regression coefficients of the composite monitors Cn. . . . . . . . . . . . . . 145

6.20 Run I systematic uncertainty estimates for the front lumi. The columns labeled

cComp and clumi contain the dominant regression coefficient for the composite

monitor and the lumi, respectively. . . . . . . . . . . . . . . . . . . . . . . . . 146

6.21 Run II systematic uncertainty estimates for the front lumi. The columns

labeled cComp and clumi contain the dominant regression coefficient for the

composite monitor and the lumi, respectively. . . . . . . . . . . . . . . . . . . 146

6.22 Luminosity monitor asymmetry results. . . . . . . . . . . . . . . . . . . . . . 149

7.1 Analyzing powers computed from simulation. The overall average AP is de-

termined by weighting the entries with the corresponding uncertainty on APV . 152

Page 24: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

1

Chapter 1

Introduction

The E158 experiment is designed to be a precision low-energy test of the electroweak theory,

complementing the substantial high-energy results obtained by the SLC and LEP colliders.

This introductory chapter provides a background to the experiment by briefly tracing the

evolution of the understanding of the weak force and its unification with the electromagnetic

interaction in the framework of the Standard Model. In addition, the methodology and

timeline of the E158 experiment are outlined.

1.1 Early Study of the Weak Force

The study of the weak force has its beginning in the year 1900 with the discovery of β

radiation by Becquerel [1]. The process was understood as the decay of an atom in state X

to another state X ′ through the emission of an electron e−:

X → X ′ + e−. (1.1)

The decay results in only two products, so conservation of energy dictates that all β decays

should emit an electron with the same energy.

In 1914, Chadwick made the surprising discovery that the energy of the particles emitted

Page 25: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

2

in β decay was not a single value, but a continuum [2]. To many scientists, the result

indicated that β decays simply did not conserve energy. Pauli remedied the situation in

1930 by hypothesizing that there was another particle ν1 emitted in β decay, along with

the electron, carrying off the missing energy [3]:

X → X ′ + e− + ν. (1.2)

The particle had to be very light and weakly interacting to have avoided detection. It was

clear that the ν particle could not interact electromagnetically but only through a new

force, called the weak force.

Fermi dubbed the new particle the “neutrino” and in 1934 incorporated it into a success-

ful theory describing β decay. Figure 1.1 depicts the diagram representing Fermi’s model.

The four particles of the decay interact at a single vertex, with a coupling strength denoted

as GF [4].

Figure 1.1: Fermi’s four-point interaction for β decay.

In 1949, it was realized that GF was identical for many different weak decay processes.

This led Lee, Yang, and Rosenbluth to postulate that all weak interactions are mediated

by a massive boson, named the W− (or its anti-particle, the W+) [5]. Figure 1.2 represents

β decay with the introduction of the W−. For energy scales below the mass of the W ,1Pauli actually named the new particle “neutron,” but that name was taken in 1932 with Chadwick’s

discovery of the neutral partner to the proton in the nucleus. Also, the particle ν is actually called theanti-neutrino ν today.

Page 26: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

3

the process depicted in Figure 1.2 is indistinguishable from Fermi’s single-vertex process in

Figure 1.1.

Figure 1.2: The W− incorporated into β decay.

1.1.1 Parity Violation

In the early 1950’s, it was observed that two particles called τ+ and θ+ appeared to be

identical in all respects, except that they decayed to states of opposite parity. Lee and

Yang speculated in 1956 that the τ and θ particles were actually the same particle, with

the weak decay mode violating parity [6]. They pointed out that while there was ample

evidence that strong and electromagnetic interactions conserve parity, there was no such

evidence for weak interactions.

The following year, an experiment by C. S. Wu with polarized 60Co demonstrated un-

equivocally that indeed the weak force did not respect parity [7]. Since that time, parity

violation experiments have been an important probe of the structure of the weak interac-

tions.

1.1.2 Electroweak Unification

In 1961, Glashow presented the first work attempting to unify the weak and electromagnetic

interactions into a single framework [8]. In addition to the weak mediators W±, the theory

predicted a weak neutral current, mediated by the Z0. In 1967, Weinberg and Salam cast

the unified electroweak theory in the form of a gauge theory with spontaneous symmetry

Page 27: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

4

breaking to explain the difference in the masses of the weak mediators and the photon [9, 10].

The Glashow-Weinberg-Salam (GWS) theory contains a free parameter, the weak mixing

angle θW , that determines the relative strengths of the electromagnetic coupling ge and the

weak couplings gW and gZ through

gW = ge

sin θWand gZ = ge

sin θW cos θW.

The theory also relates the masses of the weak mediators through

MW = MZ cos θW . (1.3)

The theory was bolstered in 1973 with the discovery of a weak neutral current event in

the Gargamelle bubble chamber at CERN [11]. They observed the interaction

νµ + e− → νµ + e−, (1.4)

which could only be mediated by the Z0 boson.

In 1978, the E122 experiment at SLAC observed the parity-violating asymmetry in

polarized electron scattering from a deuterium target. The results were in agreement with

the GWS theory, and ruled out competing models [12, 13]. The weak mixing angle was

measured to be sin2θW = 0.224 ± 0.020.

The knowledge of sin2θW and GF are sufficient to calculate the masses of the W and

Z0 particles in the GWS theory. Using the result of the E122 experiment, it was found

that the W particle should have a mass of ≈ 80 GeV/c2, while the Z0 should have a mass

of ≈ 90 GeV/c2. In 1983, the W± and Z0 were found at CERN at the predicted energies,

dramatically supporting the GWS electroweak theory [14].

Page 28: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

5

1.1.3 LEP and SLC

The SLC at SLAC and LEP at CERN were e+e− colliders that operated in the 1990’s to

test the Standard Model to high precision near the mass of the Z0. With few exceptions,

the results were in excellent agreement with the theoretical predictions. The precision of

the collider results can be appreciated by noting that the weak mixing angle was measured

to be

sin2 θW = 0.23113 ± 0.00015, (1.5)

combining all data [15].

1.2 The Role of the E158 Experiment

As with any coupling constant, renormalization causes sin2θW to run as a function of the

four-momentum transfer Q of an interaction. Figure 1.3 depicts the predicted running of the

weak mixing angle with energy. In order to fully explore the Standard Model, measurements

must be performed at several different Q values. The E158 experiment complements the

work done at the SLC and LEP by operating in an energy range far below the mass of the

Z0. Interference with the dominant electromagnetic diagrams allows low Q2 experiments

unique sensitivity to physics effects beyond the Standard Model.

1.3 E158 Overview

The E158 experiment extracts sin2θW at a Q2 of 0.026 (GeV/c)2 by measuring the parity-

violating asymmetry APV in Møller scattering. The experiment is technically challeng-

ing because the asymmetry is expected to be very small, on the order of -150 parts-per-

billion. The measurement represents the first time that parity violation has been observed

Page 29: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

6

Figure 1.3: The running of sin2θW with four-momentum transfer Q.

in electron-electron scattering.

Figure 1.4 presents a general diagram of the E158 apparatus. The experiment utilizes

the 48 GeV polarized electron beam provided by the two-mile linear accelerator at SLAC,

scattering off of unpolarized atomic electrons in a fixed target of liquid hydrogen. The

scattered flux is then integrated with the E158 calorimeter, located in the experimental hall

End Station A. The asymmetry APV is defined as the difference in the scattering rate for

each beam helicity, normalized to their sum, given by

APV =dσdΩ

R − dσdΩ

L

dσdΩ

R+ dσ

L, (1.6)

where L and R refer to the helicity of the incident beam.

Figure 1.4: General diagram of the E158 apparatus. The regions of the E158 calorimeterare named as follows: A.) In ring B.) Mid ring C.) Out ring and D.) eP detector.

Page 30: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

7

The calorimeter is divided into four annular regions, designated as In, Mid, Out, and

eP. The E158 spectrometer focuses the scattered electrons so that the inner three regions

of the calorimeter are dominated by Møller scattering events, while the outer eP ring is

dominated by electron-proton scatters. The In and Mid region together are known as the

Møller detector2, and are used for the measurement of APV .

1.4 Experiment Timeline

The data for the E158 experiment was collected over four distinct periods, designated as

Run 0 through Run 3. Each data set is analyzed independently to determine sin2θW . The

results of each Run are then combined to obtain the overall E158 result.

Table 1.1 presents the amount of data in each Run. This paper covers the combined

results of Run 1 and Run 2, representing slightly over half of the full E158 data set. The

analysis of Run 3 is still underway.

Dataset Time Total Data (pulses)Run 0 Winter 2001 Engineering RunRun 1 Spring 2002 212 millionRun 2 Fall 2002 236 millionRun 3 Summer 2003 360 million

Table 1.1: The E158 dataset boundaries.

2The Out region could have been included, but was found to be susceptible to large systematic uncer-tainties (Section 6.7.2.2).

Page 31: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

8

Chapter 2

Theory

The SLAC E158 experiment measures the parity-violating asymmetry in Møller scattering

at a Q2 five orders of magnitude below the weak scale. The measurement probes the

Standard Model at the one loop level, providing insight into the running of the electroweak

observable sin2θW .

This chapter presents the theoretical prediction for the parity-violating asymmetry APV

in Møller scattering and its relation to the weak mixing angle. Additionally, previous low

Q2 electroweak experiments and their results are described.

2.1 The E158 Experiment

The experiment utilizes a 48 GeV polarized electron beam on unpolarized atomic electrons

in a liquid hydrogen target to measure the parity-violating asymmetry in Møller scattering,

at a Q2 of 0.026 (GeV/c)2. Radiative corrections reduce the tree-level asymmetry by 40%.

The large relative size of higher-order effects allows the experiment to be a sensitive probe of

the Standard Model. Sections 2.3 and 2.4 present the theoretical calculation of the expected

asymmetry.

The effect from radiative corrections can be neatly accommodated by defining a Q2

dependent weak mixing angle. Figure 2.1 displays the running of sin2θW (Q2) as well as

Page 32: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

9

the results of several precision electroweak experiments. The NuTeV point refers to a

Figure 2.1: The running of sin2θW with Q.

neutrino experiment conducted at Fermilab [16], and the APV point represents atomic

parity violation studies performed by NIST and the University of Colorado [17]. These

experiments are described in Sections 2.2.1 and 2.2.2.

2.2 Previous Low Q2 Electroweak Measurements

It is clear from Figure 2.1 that the Z-pole experiments have measured sin2θW to high

precision, and the results are in agreement with the Standard Model. In contrast, the low

Q2 regime has been probed by only two experiments, with much less precision. The lack of

electroweak measurements for low Q2 is the primary motivation for the E158 experiment.

Both the NuTeV experiment and the APV measurements require considerable theoreti-

cal input to extract the weak mixing angle. The E158 experiment is designed to complement

these experiments by examining the comparatively clean process of Møller scattering. Also,

the energy scale is between the previous measurements, allowing the E158 experiment to

provide a unique point on the sin2θW (Q2) curve.

Page 33: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

10

2.2.1 NuTeV Experiment Overview

The NuTeV experiment compared neutrino and anti-neutrino scattering rates to determine

sin2θW [16]. The high purity neutrino beams were delivered by the Fermilab accelerator,

and cross sections were measured in a 120-foot-long steel detector. The relevant Feynman

diagrams are shown in Figures 2.2 and 2.3. The Q2 of the experiment was 20 (GeV/c)2.

Figure 2.2: Neutrino scattering diagrams.

Figure 2.3: Anti-neutrino scattering diagrams.

The weak mixing angle is extracted from the data by constructing the Paschos-Wolfenstein

ratio R− [18], defined as

R− ≡ σ(νµN → νµX) − σ(νµN → νµX)σ(νµN → µ−X) − σ(νµN → µ+X)

. (2.1)

The quantity R− is directly related to the weak mixing angle through

R− =12− sin2 θW . (2.2)

Page 34: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

11

The NuTeV Collaboration reports a value for sin2θW that is 3σ above the Standard Model

prediction.

Many attempts have been made to reconcile the NuTeV result within the framework of

the Standard Model. The primary focus has been on understanding the parton distribution

functions used in the determination of R−. Nuclear effects are complex, and it is possible

to shift the value of sin2θW based on the assumptions of these distributions [19]. The

induced shifts have not yet been found to align significantly the NuTeV result with the

Standard Model prediction. However, a recent O(α) re-analysis of deep-inelastic neutrino

scattering indicates that perhaps the theoretical uncertainties used in the extraction of

sin2 θW are enough to reconcile the 3σ shift in the NuTeV result. The work in this area is

still ongoing [20].

2.2.2 Atomic Parity Violation Overview

Electroweak experiments in atomic physics measure the perturbation of electronic orbitals

induced by Z0 exchange (Figure 2.4). The potential of the nucleus can be described as due

to the standard electric charge Z and the weak charge QW . The Q2 is very low, set by the

energy scale of the atomic orbitals.

Figure 2.4: Tree-level contributions to the nuclear potential.

The Z0 diagram produces small mixings of the unperturbed orthogonal electron orbitals.

The overlap allows otherwise forbidden transitions to occur. In principle, atomic parity

Page 35: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

12

violation experiments measure transition rates among these states to obtain QW .

At tree level, the weak charge of the nucleus is given by

QW = −N + Z(1 − 4 sin2 θW ), (2.3)

where N is the number of neutrons and Z is the number of protons [21]. Since sin2θW

is numerically very close to 0.25, Equation 2.3 essentially reduces the weak charge to the

number of neutrons N . Radiative corrections alter the tree-level prediction in Equation 2.3,

making atomic parity violation experiments sensitive to the running of the weak mixing

angle [22, 23].

Because the weak charge is proportional to the number of neutrons in the nucleus, heavy

atoms are the preferred subjects in APV experiments. However, to extract sin2θW from a

measurement, it is necessary to have a precise model of the electronic wavefunctions involved

in the transitions observed. The determination of the wavefunctions in heavy atoms is

complex and is the greatest source of uncertainty for atomic parity violation measurements.

The most precise APV experiment utilized 133Cs [17]. Cesium is a good subject because

it is a heavy atom with a single valence electron. The experiment initially reported a weak

charge that was 2.5σ below the Standard Model prediction. However, subsequent complex

electron wavefunction analyses [24, 25, 26, 27, 28, 29] have moved this value to within 0.5σ

of the theoretical prediction. The most recent result is plotted in Figure 2.1.

Page 36: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

13

2.3 E158 Experiment at Tree Level

The E158 experiment measures the parity-violating asymmetry APV in Møller scattering.

The asymmetry is defined as

APV ≡dσdΩ

R − dσdΩ

L

dσdΩ

R + dσdΩ

L, (2.4)

where dσdΩ is the differential cross section, and L and R refer to the helicity of the electron

beam. The target electrons are unpolarized. At tree level, there are four diagrams which

contribute to APV , depicted in Figure 2.5. In the limit that m2e− << Q2 << m2

Z0, the

+ Crossed Diagrams

Figure 2.5: Tree-level contributions to APV .

asymmetry is given by [30]

APVTree =−GµQ2

√2πα

1 − y

1 + y4 + (1 − y)4(1 − 4 sin2 θW ), (2.5)

where

y ≡ 1 − cos θCM2

. (2.6)

Gµ is the Fermi constant obtained from the muon lifetime formula [31], and the fine structure

constant α is roughly 1137 , as is suitable for low Q2 measurements. The θCM term refers to

the center-of-momentum scattering angle.

Because sin2θW ≈ 14 , APV is very sensitive to the weak mixing angle. The relation is

Page 37: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

14

made clear by noting

∆APV

APV=

∆ sin2 θW14 − sin2 θW

. (2.7)

The final E158 result is expected to measure APV to the precision of 12%, corresponding

to a determination of sin2θW to ±0.0014.

It is important to note that the tree-level asymmetry given in Equation 2.5 is a very

small number. Even at its maximum (y = 12 ), APVTree is only -300 ppb (parts-per-billion).

It is the smallness of this number that presents the greatest challenge to the success of the

experiment.

2.4 Radiative Corrections to APV

Because the tree-level asymmetry is suppressed by a factor of 1−4 sin2 θW , the contribution

of higher-order diagrams is effectively enhanced. This feature is responsible for the running

of the weak mixing angle in Figure 2.1, and allows the E158 experiment to be a sensitive

probe of the Standard Model.

Marciano has evaluated APV including one-loop radiative corrections and found the

corrected asymmetry to be given by [32]

APV =−ρGµQ2

√2πα

1 − y

1 + y4 + (1 − y)4×

1 − 4κ(0) sin2 θW +α

4π sin2 θW

− 3α(1 − 4 sin2 θW )32π sin2 θW cos2 θW

[1 + (1 − 4 sin2 θW )2] + F1(y,Q2) + F2(y,Q2),

(2.8)

where

sin2 θW ≡ sin2 θMSW (M2

Z0). (2.9)

Page 38: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

15

The term in the braces now has a Q2 dependence, producing the running of sin2θW depicted

in Figure 2.1. The precise definition of sin2θW (Q2) is given in Section 2.4.5.

The one-loop corrections reduce the tree-level prediction for APV by ≈ 40%. The

expected asymmetry for the E158 experiment is then roughly -180 ppb, for y = 12 . The

following sections describe the sources and sizes of the terms in Equation 2.8.

Because the effect of higher-order diagrams is large for low Q2, the E158 experiment

can be used as a sensitive probe for physics effects beyond the Standard Model at the TeV

scale. Section 7.2 describes the new physics limits that can be set with the experiment, as

well as presenting the current limits set by previous experiments.

2.4.1 γ − Z0 Mixing Diagrams

The largest one-loop corrections are contained in the κ(0) term in Equation 2.8. The

relevant processes are represented by γ − Z0 mixing diagrams and the W contribution to

the electron anapole moment, depicted in Figure 2.6. The Q2 dependence of these diagrams

is contained in F2(y,Q2).

+ Inverted + Crossed Diagrams

Figure 2.6: Photon-Z0 mixing diagrams and the W contribution to the anapole moment.

The fermionic loop in the first diagram in Figure 2.6 presents the greatest calculational

Page 39: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

16

challenge. The quark contribution cannot be evaluated perturbatively and must be deter-

mined from e+e− →hadrons experimental data. The uncertainty on κ represents the largest

contribution to the theoretical uncertainty on APV . Evaluating the diagrams, κ is found to

be

κ(0) = 1.0301 ± 0.0025. (2.10)

The 3% correction to sin2θW by κ(0) corresponds to a 37% reduction in APV . It should be

noted here that F2(12 , 0.026 (GeV/c)2) = 0.00002, a negligible contribution.

2.4.2 Heavy Box Diagrams

Box diagrams containing heavy bosons comprise the next correction to APVTree. These dia-

grams are depicted in Figure 2.7. The W diagram contributes the α4π sin2 θW

term in Equa-

+ Crossed Diagrams

Figure 2.7: Heavy boson box diagrams.

tion 2.8. It yields a 4% increase in APV over the tree-level expression. The Z diagrams

produce the −3α(1−4 sin2 θW )32π sin2 θW cos2 θW

[1 + (1 − 4 sin2 θW )2] term in Equation 2.8, resulting in a neg-

ligible 0.1% shift to the tree-level expression.

2.4.3 γ − Z0 Box Diagrams

The final contribution to the one-loop corrected value of APVTree comes from box diagrams

containing both a photon and a Z0 exchange. The relevant diagrams are depicted in Fig-

Page 40: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

17

ure 2.8. The Z0 contribution to the electron anapole moment also contributes at this level,

shown in Figure 2.9.

+ Crossed Diagrams

Figure 2.8: Photon-Z0 box diagrams.

+ Inverted + Crossed Diagrams

Figure 2.9: Anapole moment contribution from the Z0.

The contribution of these diagrams is contained in the F1(y,Q2) term in Equation 2.8.

Evaluating F1 for the values appropriate for the experiment, it is found that

F1

(12, 0.02 (GeV/c)2

)= −0.0041 ± 0.0010. (2.11)

The F1 function constitutes a 6% reduction in APVTree, largely canceling the previous W box

diagram term.

Page 41: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

18

2.4.4 ρ Term

The ρ term in Equation 2.8 follows from the convention chosen for Gµ and renormalization

of the Z amplitude [33]. The explicit form is

ρ = 1 +α

4π×

34 sin4 θW

log cos2 θW − 74 sin2 θW

+3

4 sin2 θW

m2t

m2W

+ (2.12)

3m2H

4 sin2 θWm2Z

log cos2 θW

m2Z

m2H

cos2 θW − m2H

m2Z

+1

cos2 θW

log m2H

m2Z

cos2 θW − m2H

m2Z

,

where mt refers to the mass of the top quark, and mH denotes the mass of the Standard

Model Higgs. The dependence on these masses is very slight. The other masses, mZ and

mW , refer to the standard weak gauge bosons. Assuming mt = 170 GeV/c2 and mH =

200 GeV/c2, one finds ρ = 1.00122, a totally negligible correction to the overall asymmetry.

2.4.5 Definition of sin2θW (Q2)

The precise definition of sin2θW (Q2) is a matter of convention. Often, only the terms due

to the γ-Z mixing and the W contribution to the electron anapole moment (Section 2.4.1)

are grouped into the definition of the weak mixing angle through

1 − 4 sin2 θMSW (Q2) ≡ 1 − 4κ(0) sin2 θMS

W (M2Z) + F2(Q2). (2.13)

The definition established in Equation 2.13 is typically preferred by theorists, and was used

to produce Figure 2.1.

On the other hand, experimentalists usually report an “effective” weak mixing angle.

Page 42: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

19

This amounts to defining sin2θW (Q2) so that the tree-level asymmetry formula holds, with

1 − 4 sin2 θeffW (Q2) ≡ (2.14)

ρ

1 − 4κ(0) sin2 θW +

α

4π sin2 θW

− 3α(1 − 4 sin2 θW )32π sin2 θW cos2 θW

[1 + (1 − 4 sin2 θW )2]

+F1(y,Q2) + F2(y,Q2)

Due to cancellations, the two definitions are very nearly equal for low Q2. For the parameters

of the E158 experiment, one finds sin2 θMSW = 0.2381 and sin2 θeffW = 0.2385.

Page 43: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

20

Chapter 3

E158 Beamline and BeamMonitoring

This chapter describes the major components of the E158 apparatus, from the beginning

of the accelerator up to the final collimation before the detectors. The polarized beam,

precision beam monitors, the liquid hydrogen target, and the spectrometer are covered.

3.1 Polarized Source

The helicity of the primary electron beam is controlled at the polarized source, located

upstream of the linac (Figure 3.1). The source houses a complex optical system, depicted

in Figure 3.2, which is employed to produce high beam polarization while minimizing sus-

ceptibility to helicity-correlated systematic effects [34, 35]. The following sections detail the

components of the optical system and their relevance to the E158 experiment.

Figure 3.1: Location of the polarized source room.

Page 44: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

21

Figure 3.2: E158 optics configuration at the source.

3.1.1 Laser Bench

Laser light production and pulse-shaping occur on the laser bench. The system begins with

a Flash:Ti laser which generates 12 µs pulses of linearly polarized light. A Brewster tuner

is utilized to control the wavelength of the laser, holding it to within 4 nm of the central

854 nm wavelength. At 120 Hz, the laser power is roughly 2 W.

The Slice Pockels cell is used to sample the 300 ns portion of the laser pulse with the

most favorable characteristics, balancing intensity with jitter. The Slice cell is a piezoelectric

crystal, with optical properties that are affected by the applied voltage. At zero Volts the

crystal is optically neutral, while at 3000 Volts it functions as a half-wave plate. The

Slice cell sits between two crossed linear polarizers. Biasing the Slice as a quarter-wave

plate allows the laser light to pass through both polarizers. When the cell is unbiased, an

extinction ratio of 500 is achieved.

The Tops Pockels cell is used to shape the time profile of the laser pulse, to match the

properties of the cathode. The Tops cell is placed between aligned linear polarizers and is

Page 45: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

22

pulsed at low voltage to produce mild light extinction.

3.1.2 Combiner Bench

The elements on the combiner bench are only used for diagnostics. The harmonic beam

splitter (HBS) diverts roughly 2% of the laser light to two separate beam monitors. The

spectrometer records the laser wavelength, while the photodiode is used to monitor laser

power.

3.1.3 Wall Bench

The components on the wall bench are used to define the helicity of the electron beam and

to suppress helicity-correlated beam asymmetries. The intensity asymmetry (IA) Pockels

cell is part of a feedback system designed to reduce the charge asymmetry of the electron

beam. Analogous to the Tops cell, the IA cell is operated at low voltage between aligned

linear polarizers. It is pulsed based on charge measurements performed early in the linac.

The circular polarizer (CP) and phase shifter (PS) Pockels cells are responsible for

defining the helicity of the beam that ultimately reaches the cathode. The CP cell is

pulsed at its quarter-wave voltage, converting the incoming light from linear to circular

polarization. The helicity of the light is reversed by changing the sign of the voltage bias.

The PS cell is run at lower voltages and is used primarily to correct for residual linear

polarization left by the CP cell.

The piezomirror is used to reduce helicity-correlated beam position asymmetries. Like

the IA Pockels cell, the piezomirror is controlled based on beam measurements using mon-

itors early in the linac. The voltage applied to the mirror changes the angle of reflection,

ultimately moving the laser spot on the cathode.

Page 46: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

23

The asymmetry inverter is used to combat helicity-correlated effects due to asymmetries

in the laser profile on the cathode. The inverter lenses can be moved as a unit between two

configurations. The settings complement each other, with one spatially inverting the beam

profile compared to the other. The lenses were toggled once in Run 1 and once in Run 2.

Following the wall bench, the laser light travels down the Optical Transport System

(OTS) to the cathode. The OTS is essentially a 20 m pipe, filled with nitrogen, linking the

source room with the cathode room. It contains several lenses configured to preserve the

quality of the beam.

3.1.4 Polarized Gun

The gun bench holds the final optics preceding the cathode. The lenses on the bench are

configured as a telescope, used to match the laser spot size to the dimensions of the cathode.

A mirror can be inserted after the lenses to divert the laser beam to a diagnostic target.

Optically, the target is in the same position as the real cathode. Centering light on the

target ensures that the light will be centered on the cathode when the mirror is removed.

The target is monitored remotely by a camera.

The final component on the bench is an insertable half-wave plate. It is used to combat

helicity-correlated systematic effects by reversing the helicity of the laser light defined with

the CP cell, while leaving the rest of the system unchanged. The wave plate is toggled once

every two days during E158 production running.

The cathode used for the E158 experiment is composed of a strained GaAs lattice [36].

This type of cathode has been found to provide the highest presently achievable beam

polarization along with high current. The polarization of the E158 electron beam was

measured to be ∼85% (Section 4.2), with no evidence that the cathode was charge limited.

Page 47: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

24

3.1.5 Helicity Sequence

The helicity of the electron beam is defined in sets of four pulses. The helicity of the first

pair is chosen randomly, while the second pair is the conjugate of the first. Each quad of

pulses then contains two separate sets of pulses, with the first pulse paired with the third

and the second pulse paired with the fourth. Figure 3.3 illustrates the pulse sequence.

Figure 3.3: Helicity of the E158 electron beam.

Pairs of pulses, rather than single pulses, are the fundamental unit of the experiment.

Quantities of the form

PulseR − PulseL

PulseR + PulseL(3.1)

are called asymmetries, where L and R refer to the helicity state. The results from the

detectors are reported as asymmetries. The units appropriate for the E158 experiment are

ppm or ppb, signifying parts-per-million or parts-per-billion. Quantities of the form

PulseR − PulseL (3.2)

are called differences, but may also be referred to as asymmetries. Beam position monitor

results are reported in this form.

Page 48: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

25

3.1.6 Beam Asymmetries

The polarized source is configured to decrease the size of helicity-correlated beam asymme-

tries. For a full E158 dataset, careful calibration of the positive and negative CP cell bias

voltages suppresses the expected intensity asymmetry from ∼1000 ppm down to ∼100 ppm.

The intensity asymmetry feedback then reduces the asymmetry to ∼100 ppb [35]. The large

suppression is critical for controlling systematic uncertainties because APV is only ≈ -150

ppb.

Figure 3.4 depicts the integrated charge asymmetry versus time for the toroid used for

the intensity feedback. The horizontal scale covers all of Run 1 and Run 2. The dotted line

indicates purely statistical scaling. The final asymmetry is at the level of a few hundred

ppb, as expected with the feedback asymmetry suppression.

Figure 3.4: Integrated charge asymmetry measured near the source, spanning all of Run 1and Run 2.

While the intensity feedback ensures that the charge asymmetry near the beginning of

the linac is suppressed, it is the asymmetry at the target, more than two miles away, that

is the relevant quantity for the E158 analysis. Figure 3.5 depicts the charge asymmetries

measured by the toroids just upstream of the target for both Run 1 and Run 2. The asym-

metry suppression due to the intensity feedback clearly translates into the target region,

Page 49: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

26

Run 1Parameter Integrated Asymmetry

Charge 210 ± 319 ppbEnergy -0.1 ± 1.4 keV

X -16.3 ± 5.6 nmY -3.0 ± 4.0 nm

X Angle 0.38 ± 0.23 nRY Angle 0.11 ± 0.07 nR

Run 2Parameter Integrated Asymmetry

Charge 496 ± 335 ppbE 0.9 ± 2.1 keVX 13.0 ± 6.7 nmY -15.9 ± 5.2 nm

X Angle 0.33 ± 0.22 nRY Angle 0.13 ± 0.11 nR

Table 3.1: Integrated asymmetries for Run 1 and Run 2.

although the convergence is not as strong as at the source.

Figure 3.5: Run 1 and Run 2 target region integrated charge asymmetry.

Since most beam parameters are correlated to charge, the intensity asymmetry feedback

also suppresses position asymmetries. (The position feedback with the piezomirror is also

employed, though it proved to be less effective than the IA feedback.) In the absence of

feedbacks, position asymmetries at the ∼100 µm would be expected for an E158 data set.

The feedbacks suppress the asymmetries to the level of ∼1 nm. Table 3.1 presents the

integrated asymmetries for Run 1 and Run 2.1 The asymmetry suppression observed is at

the level expected for the experiment [37].1The average asymmetry is computed by weighting the data with the resolution of the primary E158

detector. The asymmetries are then directly applicable to the analysis of the detector results.

Page 50: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

27

3.2 Beam Position Monitors

The position of the electron beam is measured with RF cavity beam position monitors

(BPMs). Figure 3.6 depicts the location of the E158 BPMs. The X and Y positions of the

beam at the target are measured with the target BPMs. The X and Y angles are computed

using the difference between the target and angle BPMs, which are separated by 40 meters.

The energy BPMs are located in an area of high dispersion so that a horizontal position

measurement actually corresponds to an energy determination. The three BPMs located

close to the source are employed for the source position feedback.

Figure 3.6: Location of E158 beam monitors.

3.2.1 BPM Operation

The beam position monitors are composed of three separate cavities [38]. Figure 3.7 is

a picture of an E158 BPM. When the electron beam traverses the device, it excites the

resonant electromagnetic modes of the cavities. The amplitude of the response is picked up

by an antenna in each cavity and read out as the signal.

The rectangular cavities are employed to measure X or Y position. The beam excites

either the TM210 or TM120, with an amplitude that is proportional to both beam position

and charge. The position cavities are then normalized to beam charge with the E158 toroids

(Section 3.3), leaving only the sensitivity to position.

Page 51: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

28

The final BPM cavity is cylindrical, and is only sensitive to beam charge. In principle,

it could be used for the normalization of the BPM position cavities. The toroids were used

instead because they have superior charge asymmetry resolution.

Figure 3.7: E158 beam position monitor.

Each E158 beam pulse is roughly 300 ns long. However, the pulse itself is composed

of smaller electron bunches, grouped at the 2856 MHz rate of the accelerator. The BPM

cavities respond to the bunches, producing a signal that oscillates at the accelerator rate.

The resonant frequency of the cavities is 2856 ± 0.3 MHz, designed to optimize the response

of the device.

The oscillation in the output signals is removed by mixing the BPM signals with an

auxiliary signal locked to the accelerator rate. The mixing electronics produce two outputs,

with amplitudes proportional to the amount of signal in phase or out of phase with the aux-

iliary signal. The mixer is tuned so that one of the output signals is maximized, minimizing

the other. A feedback system was employed to ensure that phase drifts are counteracted

and the primary signal remains a maximum.

Figure 3.8 depicts the two signals returned by the mixer for one of the BPM cavities.

The signal quickly increases with the 300 ns beam pulse and then exponentially decays.

Page 52: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

29

Figure 3.8: BPM signals after mixing.

The signals are read into custom built 16-bit VME ADCs (analog to digital converters)

as part of the data stream. High-resolution ADCs are required to match the resolution of

the BPMs.

3.2.2 BPM Performance

The BPMs are deployed in pairs or triplets (Figure 3.6) to allow for cross-checks between

devices in close proximity. The performance of a BPM pair is quantified by the distribution

of the agreement δ, defined as

δi ≡ 12(∆BPM1

i − ∆BPM2i ), (3.3)

where ∆BPMni indicates the position difference measured with BPMn for the ith pulse

pair. The width of the distribution is the pulse-pair resolution of the BPM pair. Figures 3.9

and 3.10 depict typical resolutions of the E158 BPMs.

The resolution of each of the BPMs exceeds the requirements of the E158 experiment.

This can be demonstrated by noting that the asymmetry measured with the Møller detector

Page 53: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

30

Figure 3.9: Angle and position BPM resolutions.

is corrected for beam asymmetries by

APV = ARaw −5∑

n=1

cn∆BPMn, (3.4)

where APV is the true physics asymmetry, ARaw is the measured detector asymmetry,

∆BPMn is the position difference measured with BPMn, and cn is the experimentally

determined sensitivity of the detector to beam motion. The sum covers the beam parameters

of energy, X and Y position, and X and Y angle. This method is called regression, and is

discussed in detail in Section 6.3. The resolution of the BPMs contributes to the resolution

Page 54: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

31

Figure 3.10: Energy BPM resolutions.

Parameter Detector Coefficient c BPM Resolution Resolution ContributionX -0.1 ppm/µm 2.4µm 0.2 ppmY -1.3 ppm/µm 3.6µm 4.7 ppm

Angle X -52.6 ppm/µR 0.13µR 6.8 ppmAngle Y 12.1 ppm/µR 0.15µR 1.8 ppmEnergy -15.2 ppm/MeV 1.1 MeV 16.7 ppmTotal 18.7 ppm

Table 3.2: BPM contribution to Møller detector resolution.

of the detector σDetector through

σ2Detector =

5∑n=1

(cnσBPMn )2, (3.5)

where σBPMn is the resolution of BPMn. The contributions to the resolution of the Møller

detector are presented in Table 3.2. The resolution of the detector is typically around 200

ppm, while the contribution due to BPM resolution is negligible, at less than 20 ppm.

The resolution of the BPMs is also constrained by the general goal that individual

contributions to the systematic uncertainty on APV be 5 ppb or less, over a full dataset.

The uncertainty on APV measured with the Møller detector due to the BPM corrections is

σPVA =5∑

n=1

cn < δn >, (3.6)

Page 55: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

32

Parameter Detector Coefficient c BPM Agreement δ σsystematicAPV

X 0.4 ppm/µm 1.0 ± 0.6 nm 0.4 ± 0.2 ppbY -1.2 ppm/µm 0.0 ± 1.0 nm 0.0 ± 1.2 ppb

Angle X -66.1 ppm/µR -0.07 ± 0.05 nR 4.6 ± 3.3 ppbAngle Y 7.3 ppm/µR 0.02 ± 0.03 nR 0.1 ± 0.2 ppbEnergy -25.9 ppm/MeV 0.0 ± 0.2 keV 0.0 ± 5.2 ppb

Table 3.3: Run 1 Estimate of uncertainty on APV due to BPM corrections.

where < δn > is the average BPM agreement for the nth BPM pair2. Table 3.3 details the

observed BPM agreement for Run 1 and the contribution to the uncertainty on APV [39].

The uncertainty on the agreement is proportional to the resolution of the BPMs.

Each uncertainty on σsystematicAPV is near the goal of 5 ppb or below for all beam parameters,

implying that the BPM resolution is adequate for the experiment.

3.3 Charge Monitors

The beam charge is measured by several toroids distributed along the beamline (Figure 3.6).

The toroids located near the source are used for the intensity feedback, while the toroids near

the target are used for normalizing the detector signals. Specifically, the raw asymmetry

ARaw measured with the Møller detector is

ARaw =

(MT

)R −(MT

)L(MT

)R+(MT

)L , (3.7)

where M and T refer respectively to the Møller detector and toroid signals, and L and R

refer to the helicity of the beam.2It will be demonstrated in Section 6.7.2.2 that this is a naive estimate. For the present purposes,

however, it is adequate.

Page 56: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

33

3.3.1 Toroid Overview

Each toroid is composed of an iron ring wrapped with copper wire, positioned around

ceramic portions of the beam pipe. When a charged particle passes through the ring,

a voltage is produced in the wire by inductance. The signal is then amplified and sent

through a rectifier before being read into the ADCs as the charge measurement.

3.3.2 Toroid Performance

Because the toroid signal is used for normalizing the detector signal, it is desirable that the

toroid intensity asymmetry resolution be much better than the Møller detector resolution

of 200 ppm.

The toroid resolution is determined experimentally in a fashion similar to the method

employed for the BPMs in Equation 3.3. The resolution is defined as the width of the

distribution of the agreement δi, given by

δi =1√2(A1

i −A2i ). (3.8)

The Ani terms refer to the asymmetry measured by toroid n for the ith beam pulse. The

prefactor is 1√2, instead of 1

2 used for the BPMs because only one toroid is used for the

charge normalization, instead of a pair.

Figure 3.11 presents the distribution of the agreement δ for a typical one-hour data run.

The width is 59 ppm, which is well below the 200 ppm resolution of the Møller detector.

The average resolution during the experiment was actually closer to 50 ppm, while periods

as low as 30 ppm were observed. The drifts in resolution appear to be coming from the

toroid amplifying electronics.

Page 57: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

34

Figure 3.11: Toroid resolution.

The toroids are also required to have a signal response as linear as the detectors, because

they are used for direct normalization. Section 4.1.4 demonstrates that the Møller detector

is linear to the level of roughly 99%, so it is required that the toroid be linear to that level

as well.

Each toroid mounted on the beamline has a single wire passing through its ring that can

be used for in situ calibration and linearity testing. A calibrator which can produce pulses

with better than 0.05% charge stability is used to test the response of the toroid to several

different currents. The curvature in the plot of the toroid versus the calibrator determines

the linearity.

Figure 3.12 presents the results of a toroid linearity test. The error bars on the plot,

due to calibrator fluctuations, are much smaller than the data points themselves. Also note

that the toroid is normalized so that its slope versus the calibrator is unity. The coefficient

of the second order term in the fit is quite small, indicating that the toroid is linear to

better than 99%, as required.

Page 58: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

35

Figure 3.12: Toroid signal versus calibrator signal.

3.4 Wire Array

The transverse spatial profile of the beam is monitored with a device called the wire array.

It is positioned directly upstream of the target (Figure 3.6) and can be remotely moved in

and out of the beam path. It is composed of two sets of 48 parallel wires, one horizontal and

one vertical, forming a grid through which the beam passes. The wires are composed of a

copper-beryllium alloy, with a diameter of 180 microns and a spacing of 356 microns. When

the beam passes through the grid, it causes the wires to lose electrons, producing a small

voltage which is read out as the signal. Foils near the wires are held at positive potential to

attract the liberated electrons to reduce the chance for recombination, enhancing the signal

by a factor of two. Figure 3.13 presents a typical profile measurement provided by the wire

array. The beam shape is roughly Gaussian in both axes, with an RMS of approximately 1

mm.

In addition to monitoring the shape of the beam, the wire array is used to measure the

helicity-correlated spot size asymmetry. The size and uncertainty of the spot size asymmetry

Page 59: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

36

Figure 3.13: Wire array profiles for both X and Y axes.

have implications for the systematic uncertainty on APV measured with the Møller detector.

The detector sensitivity to spot size α can be expressed as

AMeasured = APV + α∆S, (3.9)

where AMeasured is the asymmetry measured with the detector, APV is the true physics

asymmetry, and ∆S is the spot size asymmetry. The coefficient α can be experimentally

determined by comparing the widths of the distributions of the hybrid asymmetries

J+ ≡ AMeasured + η∆S (3.10)

and

J− ≡ AMeasured − η∆S. (3.11)

The η term is an arbitrary scale factor inserted for dimensional concerns, and also functions

to make the detector and wire array contributions to J± roughly the same size. In the

absence of spot size sensitivity, the width of the distributions of J+ and J− would be the

Page 60: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

37

same. The degree to which they differ determines the size of α through

α =σ2J+

− σ2J−

4ησ2∆S

. (3.12)

Figure 3.14 depicts the spot size sensitivity determined with this method for each de-

tector3, for both Run 1 and Run 2. Because the coefficients differ among the detectors,

the sensitivity is dominated by geometry rather than a common effect, such as spot size

induced target density fluctuations. In all cases, the magnitude of the sensitivity is greater

in Run 1 than Run 2, most likely due to different beam tunes.

Figure 3.14: Detector spot size correlation coefficients.

The coefficients can be combined with the helicity-correlated spot size differences mea-

sured with the wire array to produce the systematic uncertainty due to spot size on the

asymmetry measured with the detectors. Tables 3.4 and 3.5 detail the computed uncertain-

ties for Run 1 and Run 2.

The third column in the tables represents the contribution to the asymmetry measured3The luminosity monitor is a low-angle detector used for a null-asymmetry measurement. It is described

in Chapter 5.

Page 61: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

38

Average Spot size Asymmetry Systematic Uncertainty EstimateDetector α (ppm/mm2) < ∆S > (10−6mm2) α < ∆S > (ppb)Møller -61 ± 4 -0.3 ± 0.4Out 546 ± 17 5.5 ± 6.9 3.0 ± 3.8

Front Lumi 138 ± 6 0.8 ± 1.0Back Lumi 227 ± 9 1.2 ± 1.6

Table 3.4: Run 1 systematic uncertainty on APV due to spot size.

Average Spot size Asymmetry Systematic Uncertainty EstimateDetector α (ppm/mm2) < ∆S > (10−6mm2) α < ∆S > (ppb)Møller 19 ± 3 -0.2 ± 0.5Out 253 ± 32 -3.7 ± 25.2 -0.9 ± 6.4

Front Lumi 72 ± 8 -0.3 ± 1.8Back Lumi 73 ± 7 -0.3 ± 1.8

Table 3.5: Run 2 systematic uncertainty on APV due to spot size.

with the detectors due to spot size. Because α < ∆S > is consistent with zero in all cases,

the uncertainty on this number will be used as the estimate for the spot size systematic

uncertainty for the detectors. The Run 2 uncertainty is larger because the wire array was

inserted for less data compared to Run 1.

3.5 Skew Quadrupole Magnet

The emission of synchrotron radiation traversing the A-Line bend leading up to the exper-

imental hall has the effect of actually improving the quality of the beam in the horizontal

direction [40]. Higher energy electrons emit more radiation while lower energy electrons emit

less, pushing all electrons toward the average energy. Because virtually all of the bends are

in the horizontal plane, there is no enhancement for the vertical beam parameters.

The E158 experiment introduced a skew quadrupole magnet near the end of the A-

Line to mix the vertical and horizontal beam parameters. The location of the magnet is

depicted in Figure 3.15. The skew quad is a standard quadrupole magnet rotated around

the beamline by 45o. By mixing the horizontal and vertical beam parameters, the overall

Page 62: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

39

quality of the beam is improved.

Figure 3.15: Location of skew quadrupole magnet.

Figure 3.16 presents the effect of the skew quad on the detector asymmetry distributions.

The width of the distribution is the resolution of the detector. The presence of the skew

quad dramatically improves the resolution of the luminosity monitor and the Out detector,

while having little effect on the Møller detector.

Figure 3.16: Effect of the skew quad on detector resolution.

The presence of the skew quad was not found to affect the beam position jitter or energy

Page 63: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

40

jitter in any discernible way. The resolution enhancement for the Out and lumi detectors

is likely a geometric effect related to the beam divergence.

3.6 Liquid Hydrogen Target

The E158 target is a 1.5-meter-long cylinder of liquid hydrogen [41]. Hydrogen was chosen

because the background electron-proton scattering events are more easily separated and

modeled than for other targets.

The hydrogen in the target is maintained at a temperature of 20 K. The density is

0.07 g/cm3, making the target 0.17 radiation lengths. The hydrogen is continually pumped

around the target loop at 10 m/s, to minimize density changes due to heating by the primary

beam. Mesh discs are positioned in the target cell, out of the path of the beam, to induce

turbulent flow and further reduce potential density fluctuations. Figure 3.17 depicts the

components of the target loop.

Figure 3.17: E158 cryotarget loop.

The hydrogen is cooled by 14 K helium gas flowing through the heat exchanger. The

helium is provided by a refrigerator located in a building adjacent to End Station A. The loop

Page 64: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

41

itself is wrapped in 30 layers of aluminized Kapton, to minimize heating due to radiation.

The entire loop sits inside a vacuum chamber, depicted in Figure 3.18. The chamber is

large enough to allow the entire target loop to be retracted out of the beamline.

Figure 3.18: Cryotarget scattering chamber.

The electron beam deposits 700 W of power in the target for the highest current and

repetition rate used by the E158 experiment. The temperature of the target is stabilized

through an adjustable heater located at the end of the heat exchanger [42]. The target

control program automatically monitors the beam current and rate, and adjusts the heater

to maintain stable running.

The target loop contains 55 liters of hydrogen, with 25 liters in the target cell. The full

loop has an explosive yield equivalent to 8 kg of TNT, a serious safety concern. The target

loop is equipped with burst-discs which rupture when the target loop pressure exceeds safe

running conditions. The hydrogen is then directed out of the End Station roof through a

100 mm wide pipe. The target vented in this manner several times over the course of the

experiment (usually due to trouble with the refrigeration) and the safety systems worked

flawlessly. The scattering chamber is also equipped with a 160 mm wide pipe to vent the

hydrogen out the roof of ESA in the unlikely event of a rupture in the target loop itself.

This catastrophic failure mode never occurred.

Page 65: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

42

3.7 Spectrometer

The E158 spectrometer is used to separate the Møller scattered electrons from backgrounds.

It is composed of a set of three dipole magnets followed by a package of four quadrupole

magnets. The spectrometer stretches roughly half of the length of End Station A, as can

be seen in Figure 3.19.

Figure 3.19: E158 spectrometer overview.

The design of the spectrometer is closely related to the geometry of the E158 calorimeter

(Section 4.1). Figure 3.20 depicts the face of the detector.

The calorimeter is divided into four annular regions designated In, Mid, Out, and eP.

The inner three regions are dominated by Møller scattered electrons and are used in the

primary measurement of APV . The outer region is dominated by electron-proton scatters

and is used for a supplementary measurement. All four regions are collectively called the

E158 calorimeter.

3.7.1 Dipole Chicane

The primary purpose of the dipole chicane is the collimation of background bremsstrahlung

photons. For the peak beam power of 500 kW, the target produces an 85 kW photon

Page 66: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

43

Figure 3.20: Regions of the E158 calorimeter.

beam. The power level of the photon beam is too high to block with material. Instead, two

collimators are employed to keep the E158 calorimeter out of the line of sight of the target.

The unblocked photons travel with the unscattered primary beam electrons to the beam

dump out the east side of End Station A. Figure 3.21 depicts a 12 GeV scattered electron

and a target photon traversing the chicane.

Figure 3.21: E158 dipole chicane.

The photon collimation is provided by two tungsten cylinders located between the dipole

magnets. Figure 3.22 depicts the first of these collimators; the second is similar.

The chicane magnet strengths are chosen such that the integral of the transverse mag-

netic field over the path of a particle is zero, making it an achromat. Charged particles

emerge from the chicane along their initial trajectory, ensuring that the shape of the scat-

Page 67: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

44

Figure 3.22: First photon collimator.

tered electron flux is unaffected by the presence of the magnets.

3.7.2 Main Acceptance Collimator

The main acceptance collimator is located between the end of the dipole chicane and the be-

ginning of the quadrupole package. The collimator acts in conjunction with the quadrupole

magnets to provide the separation of the Møller and eP scattered electrons. Figure 3.23

depicts the collimator, looking down the beamline. The flaring of the collimator edges is

done to minimize edge scattering. The collimator is 12 cm thick of copper, followed by 3

cm of tungsten, representing forty radiation lengths of material.

The primary beam and the signal flux of the luminosity monitor pass through the

central hole of the collimator. The Møller and electron-proton scatters observed with the

E158 calorimeter pass through the outer semi-circular holes. Figure 3.24 depicts the signal

flux at the calorimeter with and without the collimator in place. Note that the shape of the

distributions is greatly influenced by the quadrupole magnets, discussed in the following

section.

The In, Mid, and Out regions of the E158 calorimeter are located between radii of

Page 68: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

45

Figure 3.23: Main acceptance collimator.

15 cm and 23.5 cm. With the collimator in place, these portions of the calorimeter are

dominated by the Møller scattered electrons between the energies of 12 GeV and 24 GeV.

The eP detector is positioned between 26.1 cm and 35 cm, where the signal is dominated

by electron-proton scatters.

3.7.3 Quadrupole Magnets

The quadrupole magnets shape the signal distribution that passes through the main ac-

ceptance collimator to separate the electron-electron and electron-proton scattering events.

Figure 3.25 presents data scans performed with the profile detector (Section 4.3). The re-

sults demonstrate the size of the signal flux at the E158 calorimeter with and without the

quadrupole magnets energized.

When the magnets are off, all of the electrons go to the eP detector. When the

quadrupoles are on, the lower energy Møller scattered electrons are focused onto the In, Mid,

and Out regions of the detector. Comparing quads-on and quads-off data with the Monte

Carlo simulation provides a powerful calibration of the model of the E158 spectrometer.

Page 69: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

46

Figure 3.24: Effect of main acceptance collimator on signal flux.

Figure 3.25: Profile scans with and without quadrupole magnets.

The simulation (Section 6.7.1) is crucial for estimation of the background electron-proton

events in the In, Mid, and Out detectors.

3.7.4 Insertable Acceptance Collimator

An additional collimator is mounted on rails next to the main acceptance collimator. It

can be remotely inserted or withdrawn when required. It functions to decrease the size

of the acceptance of the main collimator, providing better separation between the Møller

Page 70: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

47

and proton scattered electrons. The trade-off is that the amplitude of the overall signal

flux is greatly reduced. The collimator is inserted for applications that do not require high

statistics, such as polarimetry. Figure 3.26 is a photograph of the insertable collimator,

with the main acceptance collimator visible in the background.

Figure 3.26: Insertable acceptance collimator.

The large openings at the top and bottom of the insertable collimator are used for

polarimetry. The four smaller openings are used for background studies useful for fine-

tuning the Monte Carlo simulation.

Figure 3.27 depicts the change in the scattered electron profile with the insertable col-

limator in the beam. The left plot is a vertical scan, observing flux from the large lower

hole of the collimator. The separation between the Møller and proton scattered electrons is

cleaner than when the collimator is removed, shown in the plot on the right. The vertical

scale on the right plot has been normalized to match the conditions of the left plot, making

apparent the signal suppression with the insertable collimator in place.

3.7.5 eP Collimator

The asymmetry in the electron-proton flux that enters the eP detector is an order of mag-

nitude larger than APV . A simulation of showering in the E158 calorimeter demonstrates

Page 71: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

48

Figure 3.27: Profile scans with (left) and without (right) the insertable collimator.

that shower leakage from the eP detector into the other regions of the detector can be an

important effect [43]. Shower leakage accounts for 35% of the eP background in the Out

detector, with roughly 15% in the Mid and In regions.

To reduce the influence of the background, the eP collimator was installed between Run

1 and Run 2 to block completely the eP detector, as well as 75% of the Out detector. The

collimator is composed of lead, 22.86 centimeters thick. It is a “clamshell” design, with

halves that are remotely insertable and removable. Figure 3.28 presents an overhead view

of the detector area, including the eP collimator, while Figure 3.29 depicts the coverage of

this collimator on the face of the E158 calorimeter.

Though the majority of the Out detector is blocked, its signal is only degraded by 25%

due to the eP collimator. The bulk of the Out signal resides at lower radii, and the Out

ring has some shower sharing with the Mid ring. Figure 3.30 presents data from profile

scans with and without the eP collimator in place. It is clear that the collimator effectively

blocks the flux of the eP detector.

Page 72: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

49

Figure 3.28: Insertable eP collimator, top view.

3.7.6 Synchrotron Collimation and Photon Masks

A background of synchrotron photons is produced by the bends in the dipole chicane. For

the highest beam current and rate, the synchrotron radiation from the final dipole magnet

(D3) is 115 W, of which approximately 10% is directed at the E158 calorimeter [44]. The

energy in the synchrotron radiation intersecting the calorimeter is comparable to the energy

contained in the total electron signal flux. The synchrotron radiation represents a sizable

dilution to the main signal, but the primary concern is that it can introduce a helicity-

correlated asymmetry of its own, several orders of magnitude greater than APV [45].

The synchrotron background is decreased by three sets of collimators. The first is

provided by 40 radiation lengths of material in the spokes of the main acceptance collimator,

seen in Figure 3.23. The second is a similar pair of spokes, comprised of 20 radiation lengths

of tungsten, installed after the final quadrupole magnet. The final collimators are also spokes

of 20 radiation lengths of tungsten, bolted directly to the face of the detector. The final

two sets of synchrotron collimators are visible in Figure 3.31. The collimators reduce the

synchrotron signal to 0.15% of the electron signal in the E158 calorimeter.

Page 73: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

50

Figure 3.29: The hatched area represents the coverage of the eP collimator on the face ofthe E158 calorimeter.

Figure 3.30: Profile scans without (left) and with (right) the eP collimator in place.

The final set of collimators are 7 tungsten rings located inside the drift pipe upstream

of the E158 calorimeter (Figure 3.31). These are used to block photons scattered from the

photon collimators between the dipole magnets. The rings were installed after the initial

engineering run of the experiment, before the beginning of Run 1. The resolution of the

Møller detector improved from 500 ppm to 200 ppm between Run 0 and Run 1, due to the

tungsten rings blocking this background.

Page 74: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

51

Figure 3.31: Drift pipe synchrotron and photon collimators.

Page 75: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

52

Chapter 4

Detectors

The data from the E158 experiment is used to determine a single quantity: the parity-

violating asymmetry in Møller scattering. However, the measurements of many different

detectors contribute both directly and indirectly to the final result. Figure 4.1 depicts the

locations of the detectors in End Station A. Most of the devices are mounted to a movable

cart, depicted in Figure 4.2. The following sections give descriptions of the E158 detectors

and their roles for the experiment1.

Figure 4.1: Overhead view of the detector locations in End Station A.

1The luminosity monitor was the primary responsibility of the author and is covered separately in Chap-ter 5.

Page 76: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

53

Figure 4.2: Overhead view of the movable detector cart.

4.1 E158 Calorimeter

The E158 calorimeter is the primary detector of the E158 experiment [46]. It is divided

into four separate annular regions, denoted as the In ring, Mid ring, Out ring, and eP

detector. The spectrometer (Section 3.7) focuses electrons scattered from the target so that

the In, Mid, and Out regions are dominated by Møller scattering, while the eP detector is

dominated by electron-proton scatters.

The In and Mid region are collectively known as the Møller detector, and provide the

primary measurement of APV . The Out ring could also have been included, but Sec-

tion 6.7.2.2 will demonstrate that it was found to be susceptible to large systematic effects.

The eP detector supplements the Møller detector by providing information on the back-

ground electron-proton scattering events.

Page 77: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

54

4.1.1 Calorimeter Design

The calorimeter is mounted on the movable detector cart located in End Station A (Fig-

ure 4.2). It is 16 radiation lengths thick, chosen as a compromise between large signal

size and minimizing sensitivity to the pion background. The calorimeter is composed of 100

copper plates interspersed with quartz fibers. The three inner regions are 10% quartz by vol-

ume, while the eP detector is 2%. Electrons shower in the copper and produce Cherenkov

light in the quartz, which is then directed through light guides to photomultiplier tubes

(PMTs) for detection. The Cherenkov angle in quartz is close to 45, so the plates and

fibers are positioned at this angle to maximize light collection efficiency. The resulting

geometry is depicted in Figure 4.3. The In, Mid, and Out rings make up the dark inner

region while the eP detector is the light outer region.

Figure 4.3: Partially constructed E158 calorimeter.

Page 78: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

55

The quartz fibers are bundled to divide the calorimeter into the four concentric zones.

Each region is further subdivided with bundles of fibers servicing separate photomultiplier

tubes. Figure 4.4 depicts the different regions of the detector. The In and eP rings are

serviced by 10 tubes, while the Mid and Out rings have 20 tubes.

Figure 4.4: Calorimeter channel map.

To protect the PMTs from radiation damage, they are located roughly 70 cm from the

beamline. The 60 light guide periscopes that direct the Cherenkov light from the calorimeter

to the PMT locations are shown in Figure 4.5.

The PMTs are also encased by a large slab of lead for further protection. Figure 4.6

presents a diagram of the lead shielding. The PMTs are positioned in the cylinders drilled

in the back shield. The two rows of PMTs are located at radial distances of 67.3 cm and

75 cm. Over the course of the experiment, the E158 calorimeter absorbed a radiation dose

Page 79: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

56

Figure 4.5: Light guide configuration.

of approximately 500 MRad, while the PMT dose was only ∼1 Rad.

Figure 4.6: E158 calorimeter lead shielding.

4.1.2 Calorimeter Electronics

The electronics for the E158 calorimeter are depicted in Figure 4.7 [47]. The RLC circuit

is employed to increase the length of the signal from the PMTs. The longer time constant

allows the ADCs (analog-to-digital converters) to integrate for a longer period of time,

Page 80: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

57

suppressing random noise. Employing this method, it was found that the ADCs had only

one to two counts of noise, compared to the full range of 64,000 counts.

Figure 4.7: E158 calorimeter electronics diagram.

The power supplies and ADCs are located in the electronics hut, which is accessible at

all times. The hut is connected to the End Station through 200 feet of cable laid in an

underground tunnel. The amplifiers are located in the End Station near the detector to

avoid amplifying pick-up noise from the cables.

4.1.3 Calorimeter Resolution

The asymmetry resolution of the Møller detector is the dominant contributor to the overall

uncertainty on the measured value of APV . Because of the large cross section for Møller

scattering, the detector receives a signal of ≈ 20 million scattered electrons for a beam cur-

rent of 5×1011 electrons. The counting statistics contribute roughly 160 parts-per-million

(ppm) to the resolution of the detector. Additionally, common-mode electronics noise con-

tributes 110 ppm to the resolution, so that the overall resolution is near 200 ppm. Figure 4.8

depicts the Møller detector asymmetry distribution for a standard one-hour data run. The

correlation of the detector with the beam monitors has already been removed through the

regression process, covered in Section 6.3.

Page 81: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

58

Figure 4.8: Møller detector asymmetry resolution.

The Out detector receives an additional four to seven million Møller electrons (depending

on whether the insertable eP collimator is in or not). If the Out detector is included with

the In and Mid regions, the detector resolution improves slightly, as seen in Figure 4.9.

Because of its large systematic susceptibility and only marginal resolution gain, the Out

detector is not included in the measurement of APV .

Figure 4.9: Møller plus Out ring asymmetry resolution.

The eP ring is dominated by electron-proton scatters, and is used to measure the parity-

Page 82: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

59

violating asymmetry in this background. Figure 4.10 depicts the eP detector asymmetry

distribution. The resolution of the eP detector is adequate to perform its function, which

will be covered in Section 6.7.1.

Figure 4.10: eP ring asymmetry resolution.

4.1.4 Calorimeter Linearity

A non-linear response can introduce an error between the asymmetry observed with the

detector AMeasured and the physics asymmetry APV . Section 5.10.1 demonstrates that the

shift introduced by the non-linearity ε of the detector is given by

AMeasured = (1 − ε)APV − εAToroid, (4.1)

where AToroid refers to the charge asymmetry of the beam measured with a toroid.

The regression procedure is employed to reduce the detector’s sensitivity to beam pa-

rameters, including the charge asymmetry, effectively removing the εAToroid term in Equa-

tion 4.1. The remaining (1 − ε)APV term is a source of systematic uncertainty. Because

APV is ≈ -150 ppb, it is important to keep ε at the level of 1% to insure that the systematic

Page 83: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

60

Run IRing LinearityIn 0.996 ± 0.013

Mid 0.986 ± 0.010Out 0.995 ± 0.012

Run IIRing LinearityIn 0.994 ± 0.012

Mid 0.989 ± 0.009Out 1.009 ± 0.011

Table 4.1: Measured linearity of the In, Mid, and Out rings of the E158 calorimeter, forRun I and Run II.

uncertainty is only a few ppb.

The linearity of the calorimeter was constrained by comparing the observed value of

a large asymmetry measured at multiple PMT input light levels [48]. The asymmetry

was provided by the iron foil used for polarimetry, described in Section 4.2. The foil was

inserted simultaneously with the liquid hydrogen target so that the signal flux was nearly

the same as during normal production running. Filters were placed in front of some of the

PMTs, reducing the light level to 1/3 or 1/2 of the signal seen during normal running. By

comparing the asymmetry recorded by each class of PMT, it is possible to determine the

linearity. Linearity of 100% would mean that all tubes would record the same asymmetry.

Table 4.1 presents the Møller detector linearity found by this method in Run I and Run II.

The results indicate that the detector non-linearity is at the 1% level, as required.

4.2 Polarimeter

The polarimeter is used periodically to monitor the polarization of the electron beam.

Measurements are performed after each energy change and source halfwave plate toggle,

corresponding roughly to one measurement every two days. In the absence of other back-

grounds, the physics asymmetry APV is related to the asymmetry measured with the Møller

Page 84: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

61

detector AMeasured and the beam polarization P through

APV =AMeasured

P. (4.2)

The uncertainty on P translates directly into an uncertainty on APV .

4.2.1 Polarimetry Method

The beam polarization measurement is performed in a special configuration of the E158

apparatus, with the liquid hydrogen target retracted and a supermendur foil placed in

the beam [49]. The foil is magnetically polarized by a pair of Helmholtz coils which are

only energized during the measurement. The insertable acceptance collimator, discussed in

Section 3.7.4, is moved in, and the spectrometer quadrupole magnets are adjusted to provide

better separation of the Møller and electron-proton scattering events. The polarimeter

detector, described in the following section, is moved onto the Møller peak in the scattered

flux, directly in front of the Møller detector.

The polarized foil produces a helicity-correlated asymmetry in the scattering rate, due to

the magnetic dipole interaction [50]. The asymmetry Am is related to the beam polarization

Pbeam, the target polarization Pfoil, and the center-of-momentum scattering angle θCM

through

Am = cos(20)PbeamPfoil(7 + cos2 θCM ) sin2 θCM

(3 + cos2 θCM)2. (4.3)

The coefficient of cos(20) is included to account for the angle of the foil with respect to

the beam. The tree-level asymmetry calculation of Equation 4.3 is adequate because higher

order corrections are found to be well below 1% [51].

The foil is comprised of 49% Cobalt, 49% Iron, and 2% Vanadium. All three of these

Page 85: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

62

elements have two valence electrons that are polarized by the magnetic field, while the inner

electrons remain unpolarized. The valence electrons account for roughly 8% of all electrons,

so the maximum polarization of the foil is around 8%. Assuming a beam polarization of

75%, Am is expected to be approximately 4%, over five orders of magnitude larger than

APV .

Because the asymmetry is large, it can be measured in a relatively short period of

time. To avoid excessive heating of the foil, runs are taken at a beam repetition rate of

10 to 15 Hz, compared to the normal rate of 120 Hz. For each measurement, data is

taken for approximately 10 minutes, accumulating 5000 pulse pairs. Figure 4.11 presents

the measured asymmetry distribution for a typical polarimetry run. While the statistical

Figure 4.11: Polarimeter asymmetry distribution for a single run.

uncertainty on the measured asymmetry is small, the polarimetry systematic uncertainties

covered in Section 4.2.3 dominate.

Immediately after each polarization run, data is taken with the foil removed, to serve as

a background measurement. In total, each polarization measurement lasted approximately

30 minutes.

Page 86: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

63

4.2.2 Polarimeter Design

The polarimeter is mounted on the detector cart between the profile monitor and the Møller

detector, as shown in Figure 4.2. Figure 4.12 is a schematic of the detector looking down-

stream from the target.

Figure 4.12: Polarimeter position on the detector cart.

The apparatus is remotely movable in the vertical direction. When not in use, the

polarimeter is lowered to a radial distance of 50 cm so that it does not interfere with the

electron flux entering the Møller detector. During a polarimetry measurement, the device

is raised to a radial distance of 21 cm, the location with the highest rate of scattered Møller

electrons.

Signal electrons create an electromagnetic shower in the polarimeter’s quartz-tungsten

sandwich, which is depicted in Figure 4.13 [52]. The shower creates Cherenkov light which

is directed through a light guide to the photomultiplier tube that records the signal. The

sandwich plates are angled at 30, which simulations indicated was optimal for light collec-

tion. The extra tungsten plate at the front of the sandwich has only a minimal influence

Page 87: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

64

Systematic UncertaintiesItem Uncertainty

Background Subtraction 3.0%Foil Polarization 3.0%Result Spread 2.5%

Polarimeter Linearity 1.5%Foil Angle 1.0%

Foil Heating Depolarization 1.0%Levchuk Effect [54] 1.0%Analyzing Power 0.5%

Helmholtz Magnetic Field 0.2%

Table 4.2: Polarimetry uncertainties, relative to Pbeam.

on the light yield, but serves to decrease sensitivity to low-energy backgrounds.

Figure 4.13: Quartz-tungsten sandwich of polarimeter.

4.2.3 Polarimeter Systematic Uncertainties

While the data acquired for each polarization measurement provides a statistical uncertainty

well below 1%, there are many sources of systematic uncertainties also to consider. Table 4.2

itemizes each effect, which are covered in detail in Reference [53].

The contribution labeled “Result Spread” refers to the results obtained with different

target foils. Three foils of differing thickness were available on the movable target chassis:

20 µm, 50 µm, and 100 µm. Each foil gave a different result for the polarization. The reason

Page 88: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

65

for this behavior is unclear, so the spread in the results is submitted as a contribution to

the systematic uncertainty.

The quadrature sum of all the items nets a total uncertainty of 5.5%, relative to Pbeam.

The beam polarization was typically around 84%, so this amounts to an absolute uncertainty

of ±4.6% on the beam polarization.

4.2.4 Polarimeter Results

Figure 4.14 shows the polarization measurements spanning one month of Run II. The vari-

ation among the measurements is only a few percent, indicating that the beam polarization

was fairly stable. The average polarizations used for computing the physics asymmetry

were found to be

Pbeam = 84.9% ± 4.6% (4.4)

for Run I and

Pbeam = 84.4% ± 4.6% (4.5)

for Run II.

4.3 Profile Monitor

The profile monitor is used to measure the radial distribution of the scattered electrons en-

tering the Møller detector. The signal flux maps are employed to fine-tune the Monte Carlo

simulation of the E158 spectrometer, which plays a critical role in background estimation

for the Møller detector measurements.

Page 89: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

66

Figure 4.14: Run II beam polarization measurements, spanning 1 month.

4.3.1 Profile Monitor Design

Figure 4.2 shows the location of the profile monitor from overhead, while Figure 4.15 presents

a schematic drawing of the device. The profile monitor is comprised of two of pairs movable

Cherenkov detectors mounted to a rotatable annulus. The detectors can be moved radially

from the beam pipe out to 50 cm, fully covering the extent of the E158 calorimeter. The

annulus can be rotated over 180, allowing radial scans at any angle. The scanners are

cross-calibrated by comparing data taken 180 apart.

The Cherenkov detectors are mounted in pairs for background subtraction [55]. The

signal from the front scanner is weighted and subtracted from the back scanner to produce

a corrected signal. The front scanner utilizes only a quartz block, while the back scanner is

also equipped with a tungsten preradiator. Figure 4.16 depicts one of the back scanners.

The shutter mounted in front of the PMT can be opened and closed remotely, for

background studies. Additionally, the outer tube of the scanner is rotatable so that the

tungsten preradiator can be removed for additional background measurements.

Page 90: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

67

Figure 4.15: Profile monitor schematic.

4.3.2 Profile Scan Results

A profile scan consists of multiple radial scans over the full azimuth, in increments of 20. A

full scan required roughly 30 minutes of beam time. Figure 4.17 shows a typical radial scan

at one particular angle. The Møller peak is at the left, and the electron-proton scattering

peak is at the right. The use of profile scans for background subtraction is discussed in

Section 6.7.1.

4.4 Pion Detector

Pion production in the target contributes a small background to the Møller scattering

process. The parity-violating asymmetry APV is related to the asymmetry obtained from

Page 91: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

68

Figure 4.16: Quartz scanner schematic.

Figure 4.17: Radial scan produced with the profile monitor.

the Møller detector AMeasured and the pion background through

APV =AMeasured − εNπ

NeeAπ

1 − εNπNee

Aπ. (4.6)

The term Nπ/Nee is the ratio of the number of pions to electrons passing through the

Møller detector. The ε term refers to the average signal size produced by a pion in the

Møller detector, compared to the average signal produced by an electron. The term Aπ

represents the parity-violating asymmetry of the pion background. The pion detector is

utilized to determine Aπ, as well as the ratio Nπ/Nee, while the remaining unknown ε must

Page 92: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

69

be found from simulations of the Møller detector [56].

Pions are produced through real and virtual photoproduction, as well as deep inelastic

scattering processes. Calculations indicate that the pion background passing through the

Møller detector is likely to have an asymmetry at the level of 1 ppm, an order of magnitude

higher than the Møller scattering asymmetry APV [37]. However, it is also found that

the ratio Nπ/Nee is expected to be less than 1%. Sensitivity to the background is further

suppressed because the Møller detector is less sensitive to pions than electrons. Simulations

show that ε is 0.22 ± 0.15. Combining these factors, it is expected that the pion asymmetry

correction to AMeasured is less than 4 ppb, with a dilution factor less than 0.004.

4.4.1 Pion Detector Design

The pion detector is located directly behind the E158 calorimeter, shown in Figure 4.2. The

lead PMT shield and active copper region of the Møller detector constitute 60 radiation

lengths of material in front of the pion detector. The material ensures that the signal in the

pion detector is dominated by pions, with little contamination from scattered electrons.

The detector is composed of 10 individual photomultiplier tubes, mounted symmetrically

around the beam pipe. A quartz block is positioned in front of each tube, defining an active

region for the detector between 15 cm and 23.5 cm, measured radially from beam center.

This region exactly matches the extent of the three inner rings of the E158 calorimeter so

that the two detectors observe the same pion flux. The tubes are mounted at 45 relative

to the beam, to maximize Cherenkov light collection efficiency. Figure 4.18 depicts the pion

detector looking downstream from the target.

Page 93: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

70

Figure 4.18: Pion detector layout.

4.4.2 Pion Detector Resolution

The required asymmetry resolution for the pion detector is set by the general goal that indi-

vidual systematic uncertainties be kept below the level of 5 ppb. Because the correction due

to the background pion asymmetry is expected to be less than 5 ppb for the Møller detector,

it is acceptable to measure the pion asymmetry to ± 100% of its value, corresponding to

a precision of ≈ 1 ppm. Since Run I and Run II are each composed of roughly 80 million

pulse-pairs of data, the RMS of the pion detector asymmetry distribution is required to be

less than 1 ppm×√8 × 107 ≈ 0.01.

Figure 4.19 presents the pion asymmetry distribution for a typical one-hour data run.

The RMS of the distribution is a factor of two smaller than the required 0.01, indicating

that the pion detector design is adequate for the experiment.

It should be noted that the resolution is 25 times less sensitive than the Møller detector.

Because of its relative insensitivity, it was found that it was unnecessary to use the regression

procedure to remove the correlation of the pion detector asymmetry to beam asymmetries:

all of regression correlation coefficients were consistent with zero.

Page 94: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

71

Figure 4.19: Pion detector asymmetry distribution, covering one run.

4.4.3 Pion Detector Electromagnetic Background

Though the pion detector is positioned behind 60 radiation lengths of material, it is still

expected that Møller scattered electrons will contribute a small dilution to the pion signal.

To measure the amount of contamination, special runs were taken with the polarimetry

foil included with the liquid hydrogen target. The runs have a large, quickly measured,

asymmetry. For these data, the Møller detector recorded an asymmetry of 522 ppm (with

negligible uncertainty) while the pion detector measured an asymmetry of 62 ppm ± 11

ppm. Therefore, it is demonstrated that the electromagnetic contamination in the pion

detector is only 13.5% ± 2.7%, which is small. The remaining signal in the pion detector is

assumed to be pions.

4.4.4 Ratio of Nπ to Nee

The ratio Nπ/Nee was determined from two special runs. First, one of the pion detector

tubes was placed in front of the Møller detector. Data was then taken using only the

polarimetry foil as a target. The foil was unpolarized: it was used simply as a thin target,

Page 95: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

72

so that the pion tube would not saturate. The tube was then moved back behind the

Møller detector, and normal data was taken with the liquid hydrogen target. The signal

attenuation provided by the Møller detector is related by

fAttenuation =PMTBackLH2

PMTFrontFoil

× MøllerFoil

MøllerLH2, (4.7)

where the term fAttenuation designates the ratio of the signal flux at the front of the Møller

detector to the signal at the front of the pion detector. The PMT terms refer to the

signal size from the pion tube, while the Møller terms designate the signal size found by

combining all of the tubes from the Mid ring of the Møller detector. The Møller ratio is

included as a normalization, to account for the use of the different targets. It was found

that fAttenuation = 2.003 × 10−4 ± 0.346 × 10−4.

The attenuation number is compared with a GEANT3 simulation to extract the ratio

Nπ/Nee [57]. The simulation is used to determine the value of

fAttenuation =CBackee + Nπ

NeeCBackπ

CFrontee + NπNee

CFrontπ

. (4.8)

Here the C terms refer to the size of the Cherenkov response of the pion detector. The

Back and Front labels indicate the location of the pion tube, while ee and π refer to Møller

electrons and pions, respectively.

Comparing the observed fAttenuation with the simulation yields

Nee= 0.0063 ± 0.0021. (4.9)

The uncertainty is composed of the statistics of the real measurement, as well as a conser-

Page 96: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

73

Pion Asymmetry AπRun I -1.74 ± 0.46 ppmRun II -0.36 ± 0.48 ppm

Table 4.3: Pion detector asymmetry results.

vative systematic uncertainty estimate found by varying the simulation parameters.

4.4.5 Pion Detector Results

Table 4.3 presents the asymmetries measured with the pion detector for Run I and Run II.

The results for the two runs differ by three standard deviations. The cause for this effect

can be understood by noting that the signal in the pion detector is reduced by 40% when

the insertable eP collimator (Section 3.7.5) is in place. This indicates a sizable signal due to

punchthrough from the eP detector directly into the PMTs of the pion detector. Because

the asymmetry measured with the eP detector was ≈ 1.5 ppm, the contamination of the

pion detector in Run I renders it unusable. Therefore, it was decided that the Run II result

would also be employed for Run I, with the Run I result disregarded. Using Aπ, Nπ/Nee,

and ε, the correction applied to Run I and Run II is -0.5 ± 0.8 ppb, with a dilution factor

εNπNee

of 0.0014 ± 0.0011.

4.5 Synchrotron Light Monitor

The Synchrotron Light Monitor (SLM) detects synchrotron radiation produced in the A-

Line, the curved beamline connecting the linac to End Station A [58]. The general location

of the detector is depicted in Figure 4.20.

The device is used to quantify the vertical polarization of the beam by measuring the

helicity-correlated asymmetry in the synchrotron radiation. The asymmetry has relevance

to the Møller detector, because it also presents itself in the synchrotron radiation background

Page 97: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

74

Figure 4.20: SLM location in the A-Line bend.

produced in the dipole magnets in End Station A.

4.5.1 Synchrotron Asymmetry

The power emitted by an electron beam traversing a magnet has an asymmetry related to

the amount of spin polarization in the direction of the magnetic field [45, 59]. The power

in a synchrotron beam is given by

P = P0

[1 −

(55√

324

+ η

], (4.10)

where P0 refers to the classically calculated synchrotron radiation power, and η = ±1 refers

to the electron spin oriented parallel or anti-parallel to the magnetic field. The χ term

reflects the properties of the applied magnetic field through

χ =3hγ2

2mecρ, (4.11)

with ρ being the bend radius. The asymmetry in the synchrotron radiation power is found

to equal χ through

ASR ≡ P+ − P−

P+ − P− = χ. (4.12)

The final dipole magnet in the E158 spectrometer, D3, produced the vast majority of

Page 98: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

75

the synchrotron radiation background for the Møller detector. Using the values for this

magnet, it is found that the power asymmetry in the synchrotron radiation is ≈ 35 ppm

for a 100% vertically polarized electron beam [44].

If the Møller detector were truly a calorimeter for photons over the energy range in the

synchrotron radiation, the analyzing power for vertical polarization would be χ. However, it

is likely that the detector has a non-linear response to synchrotron photons in the relevant

energy ranges. Simulations indicate that the energy weighted asymmetry for the Møller

detector would be closer to 65 ppm, rather than the real asymmetry of 35 ppm [60]. To be

conservative, it will be assumed that the Møller detector has an analyzing power of 60 ppm

± 30 ppm for vertical beam polarization.

By comparing the Møller detector signal sizes for empty target runs to normal runs

with the liquid hydrogen target, it was found that synchrotron radiation comprises 0.15%

± 0.05% of the total Møller detector signal. Combining this factor with the synchrotron

asymmetry and an expected vertical beam polarization of ∼1% reduces the effect of the

background to a few ppb.

The asymmetry measured with the SLM is used to place bounds on the amount of verti-

cal polarization in the electron beam, which in turn constrains the effect of the synchrotron

radiation produced in the D3 magnet on the Møller detector.

4.5.2 SLM Design

Figure 4.21 presents a schematic of the SLM. The synchrotron radiation in the SLM region

exits through a 1 cm thick aluminum flange. Then, a 1 mm thick layer of lead is used

to convert the photons into an electron shower. The shower then traverses a quartz bar,

producing Cherenkov light. The light is directed by a mirror to three photodiodes, which

Page 99: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

76

deliver the SLM signal to the ADCs. Photodiodes have low resistance to radiation, so the

mirrored box is heavily shielded with lead.

Figure 4.21: SLM layout.

4.5.3 SLM Resolution

The synchrotron radiation background affects the Møller asymmetry in a manner analogous

to the pion background given in Equation 4.6:

APV =AMeasured − fASynch

1 − f. (4.13)

The term f is the dilution factor, while ASynch refers to the asymmetry registered in the

Møller detector due to the synchrotron radiation produced in the D3 magnet of the spec-

trometer.

The value for fASynch is given by

fASynch = (0.0015 ± 0.0005) × (60 ppm ± 30 ppm) × Py, (4.14)

where the first term is the dilution factor f , the second is the analyzing power of the

Page 100: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

77

Møller detector, and Py is the vertical beam polarization. As with the pion background,

it is desirable to keep the uncertainty due to the synchrotron background below 5 ppb.

Therefore, the SLM is required to measure the vertical beam polarization to roughly 5%.

The analyzing power of the SLM is found from simulations to be 60 ppm ± 30 ppm

(coincidentally the same as the Møller detector). Therefore, over the course of the typical

data set of 80 million pulse pairs, the SLM must measure its asymmetry to the level of 2

ppm, corresponding to an asymmetry resolution of 0.03.

It was found that combining the three SLM channels actually did not significantly in-

crease the detector resolution. The dominant noise source for the detector must be common

to all three photodiodes. Therefore, it was decided to use the single SLM channel (diode 2)

with the best resolution for the SLM result. Figure 4.22 depicts the asymmetry distribution

of the SLM data for a typical one-hour data run.

Figure 4.22: Regressed SLM channel 2 asymmetry distribution, for one run.

With an RMS of 840 ppm, the SLM greatly exceeds the resolution requirements of the

E158 experiment. Unlike the pion detector, the resolution is sufficient that regression against

beam monitors is useful. Like the Møller detector and the luminosity monitor, the SLM is

Page 101: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

78

Vertical Polarization (%)45 GeV 48 GeV

Run I -1.7 ± 0.9 1.7 ± 0.9Run II -1.7 ± 0.9 3.1 ± 1.6

Table 4.4: Vertical beam polarization at the target.

regressed against beam charge, energy, X and Y position, and X and Y angle. Regression

produces a factor of two improvement in the resolution, resulting in the distribution seen

in Figure 4.22.

4.5.4 SLM Results

The SLM asymmetry results are presented in Figure 4.23. The data is split into four separate

chunks, based on energy and source halfwave plate (HW) state. The cancellation between

the energy states is good, effectively reducing the systematic effect due to synchrotron

radiation for the Møller detector.

Figure 4.23: SLM asymmetry results for Run I and Run II.

Table 4.4 presents vertical beam polariation results for both Run I and Run II, computed

using the SLM asymmetry. The amount of vertical polarization is ∼1%, consistent with

expectations.

Using the Møller detector analyzing power and the synchrotron dilution factor, the

Page 102: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

79

Diluted Synchrotron Asymmetry (ppb)45 GeV 48 GeV

Run I -1.6 ± 1.2 1.5 ± 1.2Run II -1.6 ± 1.2 2.8 ± 2.2

Table 4.5: Synchrotron asymmetry correction for the Møller detector.

diluted asymmetry contribution to the Møller detector asymmetry fA can be calculated.

Table 4.5 details the results, which are found to be at a manageable level.

Page 103: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

80

Chapter 5

Luminosity Monitor

The luminosity monitor is designed to perform measurements complementary to the Møller

detector. The primary function of the device is to provide a statistically significant null-

asymmetry measurement by observing extremely forward angle Møller and Mott scattering.

It fills the secondary role of determining the level of density fluctuations in the liquid

hydrogen target. The detector is also used to monitor the noise properties of the beam

because of its large sensitivity to beam motion. All of these functions serve to ensure the

data quality that is used in producing the final physics result.

5.1 Detector Signal

The luminosity monitor, henceforth to be called simply the “lumi,” is positioned 70 meters

downstream of the target, at an angle of one milliradian (Figure 5.1). At this location,

about 70% of the scattered electrons that hit the detector have energies greater than 40

GeV. The average Q2 of the signal is 0.003 (GeV/c)2, an order of magnitude lower than for

the signal observed with the Møller detector [61].

The signal rate in the detector is high, with approximately 3×108 scattered electrons for

the nominal beam current of 5×1011 electrons per pulse. It is because of the large signal in

this region that the lumi is able to provide an asymmetry measurement that has statistical

Page 104: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

81

Figure 5.1: Layout of End Station A for experiment E158.

significance on the level of the result obtained by the Møller detector. Figure 5.2 depicts a

Monte Carlo simulation of the various components of the lumi signal flux [62].

Figure 5.2: Components of lumi signal at face of detector.

The signal is dominated by Møller and Mott scattering from the target. The contribution

labeled eA refers to scattering from components in the experimental apparatus other than

the target. The largest contributor to this background is the aluminum end windows of

the target cell. The final component is a small contribution from inelastic electron-proton

Page 105: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

82

scattering in the target.

The overall expected asymmetry Alumi is the sum of the individual signal component

asymmetries Ai, weighted by their signal size Si through

Alumi =∑i SiAi∑i Si

. (5.1)

Figure 5.3 depicts the asymmetries of each of the lumi signal components, multiplied by the

ratio of its signal size to the total signal. (Note that the sign of the eA asymmetry has been

reversed.) Combining each component yields an expected asymmetry of -15 ppb ± 5 ppb

Figure 5.3: Contributions to lumi asymmetry.

for the lumi. For comparison, the Møller detector is expected to observe an asymmetry of

approximately -150 ppb, a full order of magnitude larger.

Over the course of the experiment, the Møller detector is expected to observe a non-zero

asymmetry with a significance of 10 standard deviations. The lumi is expected to measure

Page 106: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

83

essentially zero asymmetry, with the same statistical significance. As a result, the lumi can

be used to augment the Møller detector as a sensitive monitor for false asymmetries.

5.2 Synchrotron Radiation Background

The E158 spectrometer, which is discussed in Section 3.7, has three dipole magnets used to

direct the primary beam around the collimation for photons from the target. The deflected

beam produces synchrotron radiation, which in principle could be a large background for the

lumi. The amount of power in the synchrotron beam is calculated to be about 0.02% of the

total beam power [44]. For nominal data runs, this amounts to approximately 115 Watts.

Simulations show that about 15% of this power is directed at the luminosity monitor [63].

Figure 5.4 displays a simulation of the distribution of background photons at the face of

the lumi. The dark horizontal band is the region dominated by synchrotron radiation. The

density of photons is weighted predominantly to the right side of the plot because the bend

of the final dipole is to the left. The sensitive region of the detector is superimposed as the

area between the black circles. The empty central region was omitted from the simulation

for simplicity.

The energy absorbed from actual signal electrons is expected to be about 150 Watts,

meaning that synchrotron radiation is roughly 10% of the energy of the lumi signal at the

face of the detector [64]. By itself this is a large dilution factor, but the main issue of concern

is that synchrotron radiation can have helicity-correlated asymmetries of its own [45, 59, 65,

66]. Asymmetries up to the level of 600 ppb could be generated in the synchrotron spectrum

from the E158 spectrometer. Because the expected physics asymmetry of the lumi is only

15 ppb, it is desirable to limit the detector sensitivity to synchrotron radiation to less

than 1% of the total signal. As will be demonstrated in Section 5.9, this was successfully

Page 107: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

84

Figure 5.4: GEANT simulation of photon flux at the lumi.

incorporated into the design and renders the synchrotron background contribution small.

5.3 Detector Design

The lumi is an ionization detector, comprised of 16 individual channels. These channels are

grouped into two separate, full-azimuth rings. The upstream ring is identified as the “front

lumi” and the downstream ring as the “back lumi.” The front lumi is positioned behind

seven radiation lengths of aluminum showering material. Directly behind the front lumi is

an additional four radiation lengths of aluminum, followed by the back lumi. Figure 5.5

depicts the layout of the full luminosity monitor.

The aluminum between the two sets of rings ensures that the back lumi will have a

smaller signal than the front lumi. An EGS4 simulation indicates that the reduction in

signal should be approximately a factor of three between the two rings [64]. The smaller

signal allows the back lumi to serve as a cross-check on the front ring for systematic effects

Page 108: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

85

Figure 5.5: A. Front view of one full lumi ring, with sensitivity between 7 and 10 cm. B.Side view, depicting the two lumi rings and the aluminum showering material.

related to signal size.

Each chamber houses a package of 11 parallel plates (Figure 5.6), positioned transverse

to the incoming signal electrons. A bias of 100 V is applied to alternate plates, in order

to produce an electric field between each plate pair. As an ionizing particle traverses the

chamber, it produces electron-ion pairs, which are collected on the plates as the signal.

Figure 5.6: Individual chamber design, with signal plates shaded.

The chambers are filled with nitrogen gas, at a pressure of just slightly over one at-

mosphere. The overfilling is done to ensure that the amount of oxygen in the chambers

is minimized, since it has the property that it can capture signal electrons before they

are collected on the plates. The gas is continually flowed through the chambers so that

Page 109: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

86

contaminants due to radiation damage are not allowed to accumulate.

An electron passing through the lumi parallel to the beamline encounters one centimeter

of nitrogen. It is expected that a single electron will produce ≈ 60 ionization pairs for this

amount of gas [67].

The aluminum in front of the detector rings performs the dual role of showering the main

signal and attenuating synchrotron radiation. A simulation shows that the front lumi ring

is positioned very close to the shower maximum, with a predicted multiplicity of roughly

100 [64]. It is also expected that the synchrotron radiation background is suppressed below

the level of 1% of the signal at this depth.

The large scattering rate, coupled with the gain factors from the showering in the alu-

minum and the ionization trail in the nitrogen gas, allows the chambers themselves to

produce signals on the level of several volts. The size of the output signal means that no

additional amplification electronics are required, which considerably simplifies the detector

design.

The signals travel along 100 feet of coaxial cables from the chambers to the ADCs, where

they are read in differentially (after the bias voltage is filtered out of the signal cable). By

subtracting the two signals from each chamber, noise pick-up from the transmission cables

is suppressed. Figure 5.7 is a schematic of the lumi electronics setup.

5.4 Lumi Signals

Because each pair of plates is capacitively coupled, each chamber actually produces two

signals, of opposite sign. Figures 5.8 and 5.9 display oscilloscope traces of these signals, the

first from a front lumi chamber and the second from a back lumi chamber.

The time structure of the beam provided by the accelerator can be approximated as a

Page 110: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

87

Figure 5.7: Lumi electronics setup.

Figure 5.8: Front lumi signal traces.

square pulse with a duration of 300 ns. As can be seen from the scope traces in the figures,

the lumi signal is of similar length, indicating that electron collection occurs quickly. Also

note that the front lumi has a signal three times larger in magnitude than the back lumi,

in agreement with simulations.

The fast signal observed on the scope is composed of the electrons from ionization pairs.

The ions arrive much later than their electron counterparts due to their much higher mass.

In fact, the ion signal is stretched out to such an extent that it is all but invisible on

the oscilloscope traces. Nevertheless, it is statistically beneficial to collect the signal due

Page 111: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

88

Figure 5.9: Back lumi signal traces.

to the ions. By extending the integration time of the ADCs from two microseconds to

30 microseconds, thereby collecting more ions, it is observed that the detector asymmetry

resolution improves by 50%.

5.5 Gas System

The gas system (Figure 5.10) for the lumi was designed to be very simple. The nitrogen

for the entire detector is supplied by a single bottle, located outside of End Station A. The

flow rates of the individual chambers are balanced by adjusting needle valves located on the

the output lines. During operation, the gas flow is monitored by microphones positioned

to detect the sound of nitrogen bubbles escaping through silocone oil at the end of the

gas lines. The flow rate was extremely low, and the gas bottle was changed only once per

month. The chambers are electrically isolated from each other by inserting non-conductive

tubing at the entrance and exit ports of the copper gas lines.

Page 112: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

89

Figure 5.10: Gas system configuration for lumi.

5.6 Asymmetry Resolution

The asymmetry measured with the lumi ARaw is defined as

ARaw =

(ST

)R −(ST

)L(ST

)R+(ST

)L , (5.2)

where S refers to the detector signal, L and R are the beam helicity, and T refers to the

beam charge measured by a toroid just upstream of the target.

To measure the true physics asymmetry, however, effects due to beam motion must

be removed. This is achieved through measuring the correlation between the detector

asymmetry and the various beam monitors, and then removing this correlation. This process

is known as beam regression, and will be discussed in detail in the section concerning data

analysis (Section 6.3). Regression covers six parameters, removing the detector correlations

with the X and Y beam positions, dX and dY beam angles, the beam energy E, and the

beam charge Q.

Page 113: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

90

The asymmetry is corrected on a pulse-pair basis through

APV = A−6∑

n=1

cn(bRn − bLn), (5.3)

where the sum covers the six regression parameters, the constants cn are the experimentally

determined correlation coefficients, and the bn terms are the values given by the BPMs. In

the case of the charge correction, the difference in Equation 5.3 becomes an asymmetry

measured with a toroid.

Figures 5.11 and 5.12 display the correlation coefficients for the position and angle

motion obtained from a typical run, for each lumi chamber. The large azimuthal dependence

of the coefficients is evident in the sinusoidal pattern in the plots. The numbering scheme

for the chambers is the same as found in Figure 5.5.

Figure 5.11: X and Y correlation coefficients. The averages are shown as straight lines.

The asymmetry for the whole detector is the average of the asymmetries found by the

individual chambers, given by

APVTotal =1N

N∑m=1

APVm . (5.4)

Page 114: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

91

Figure 5.12: dX and dY correlation coefficients The averages are shown as straight lines.

Here the sum is over the individual lumi channels, and N refers to the total number of

channels used. Since the front lumi and back lumi are treated as separate detectors, this

number is usually eight.

Substituting Equation 5.3 into Equation 5.4 for each chamber yields the expression for

the whole detector through

APVTotal =1N

N∑m=1

(ARawm −6∑

n=1

cmn (bRn − bLn)), (5.5)

which can be written as

APVTotal =< ARaw > −6∑

n=1

< cn > (bRn − bLn). (5.6)

The variable m is summed over the lumi channels, and n sums over the beam monitors.

The form of the asymmetry given in Equation 5.6 is exactly the same as in Equation 5.3,

except that the terms are replaced by averages over the chambers. In particular, it can be

seen that the correlation coefficient relevant for the detector as a whole is the average of

Page 115: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

92

the individual chamber correlation coefficients.

The average of the channel coefficients is plotted as a solid line in Figures 5.11 and 5.12.

The average is always much less than the amplitude seen in the coefficients of the individual

chambers, indicating that the sensitivity of the detector as a whole to beam motions is

much less than the individual chamber sensitivities. The suppression is so large because

lumi chambers are automatically gain matched by design.

The RMS of the distribution of APV is the resolution of the detector. For a given set

of data, the resolution determines how well the mean asymmetry is known. The Møller

detector typically has an RMS of 200 ppm. For the lumi to serve as a useful cross-check

on the main detector, then, it must also have an asymmetry resolution of at least 200

ppm. Figure 5.13 demonstrates the level of improvement in resolution obtained by removing

detector correlations to the beam through regression. The RMS of the corrected distribution

is roughly 100 ppm, which is at a level that meets experimental goals for the lumi design.

Figure 5.13: Left: Raw lumi asymmetry distribution. Right: Regression-corrected lumiasymmetry distribution.

Page 116: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

93

5.7 Resolution Contributions

In order to evaluate the performance of the lumi, it is instructive to itemize the contributions

to the overall asymmetry resolution. Mathematically, it can be shown that the expected

resolution is given by

σ2APV =

12N

(1+∆EE

2

)+σ2Boiling+σ

2Pedestal+(1+cQ)2σ2

ToroidResolution+5∑

n=1

c2nσ2BPMResolution.

(5.7)

The term σAPV refers to the statistical width of the asymmetry distribution. The number

of scattered electrons entering the detector is given by N , so the 12N term is a reflection

of counting statistics. The statistics are modified by the energy resolution of the detector,

represented by the ∆EE term. The next term, σBoiling, refers to contribution due to density

fluctuations in the liquid hydrogen target. The term σPedestal represents all types of elec-

tronics noise. The final two terms involve the resolutions of the toroid and beam position

monitors, indicated by σToroidResolution and σBPMResolution, respectively. The beam moni-

tors affect the lumi asymmetry resolution because they are used in the regression process,

discussed in the previous section. The degree to which the beam monitors contribute is

related by the experimentally determined correlation coefficients cn. The toroid term has a

slightly different form than that of the BPMs because it is used for both normalization and

as a regression parameter.

Each item in Equation 5.7 is separately known. The beam monitor resolutions can be

obtained directly from the data, as discussed in Section 3.2.2. Zero current beam pulses

occur at a rate of 0.5 Hz during normal running, allowing σPedestal to be measured reasonably

well. The contribution due to counting statistics is obtained from a Monte Carlo simulation.

The energy resolution of 50% is obtained from a simulation of the detector response. The

Page 117: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

94

Luminosity Monitor Resolution ContributionsParameter Coefficient BPM Resolution Contribution (ppm)

Q -0.097 ppm/ppm 28.5 ppm 26E -3.17 ppm/MeV 2.0 MeV 6X -5.36 ppm/µm 3.3 µm 18Y -0.52 ppm/µm 4.5 µm 2dX -11.48 ppm/µR 0.1 µR 1dY 4.81 ppm/µR 0.2 µR 1

Counting Statistics 45Energy Resolution 20

Boiling Noise* 50Pedestal Noise 60

Total Calculated: 98Observed: 104

Table 5.1: Contributions to the luminosity monitor asymmetry resolution. *Target boilingis covered in Section 5.8.

energy resolution is low because many of the showers are not contained by the detector.

This leaves the term due to boiling noise, which is calculated using the method described in

Section 5.8. Combining these contributions for a run provides a full resolution accounting for

the lumi. Table 5.1 presents the resolution contribution itemization for a typical one-hour

data run.

The dominant contributions to the resolution are electronics noise, boiling noise, and

counting statistics. The remaining contributions due to beam monitor resolution are small

by comparison. The observed resolution of 104 ppm compares well with the calculated value

of 98 ppm, indicating that all major noise sources are known. The resolution exceeds the

200 ppm level required by the experiment.

5.8 Target Boiling

As discussed in Section 3.6, heating from the primary beam can cause density fluctuations

in the liquid hydrogen target. Though it is not technically “boiling” that is occurring, this

is still the standard name that has attached itself to this issue. The amount of boiling is

Page 118: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

95

important to the experiment because it directly translates to noise for the detectors. Also,

if there exists an asymmetry in the density fluctuations, a false asymmetry measured by

the detectors could result.

In principle, once the lumi and Møller detector asymmetries are corrected for beam mo-

tion, the only correlation between them is in target density fluctuations and the resolutions

of the beam monitors used for the corrections. The latter contribution can be obtained

from the data, so by observing the level of common-mode noise between the two detectors,

the amount of boiling noise can be extracted. The method employed [42] begins by defining

two composite asymmetries through

J±≡AMøller±ALumi. (5.8)

Here A refers to the regression corrected asymmetries of the Møller detector and the lumi.

If the two detectors were completely uncorrelated, the RMS of the distributions of J+ and

J− would be the same. The amount that the two differ yields the common-mode noise

between the two detectors. Accounting for the noise due to beam monitor resolutions, the

amount of boiling noise is given by

σ2Boiling =

σ2J+

− σ2J−

4− (1 + cMQ + cLQ)σ2

ToroidResolution −5∑

n=1

cMn cLnσ

2BPMResolution. (5.9)

As with the preceding section, the cn terms refer to the beam correction coefficients. The

superscripts M and L refer to the Møller detector and the lumi, respectively. Similar

to Equation 5.7, the toroid term differs from the BPM terms because it is used for both

normalization and as a regression parameter.

To determine how much of this “boiling” noise is actually due to target density fluc-

Page 119: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

96

Run Label Spot Size(mm x mm) Pump Speed(speed/nominal)1 1.5 x 1.5 12 1.0 x 1.0 13 1.0 x 1.0 14 1.0 x 1.0 15 1.5 x 1.5 2/36 1.0 x 1.0 2/37 1.0 x 1.0 2/38 1.0 x 1.0 2/39 1.5 x 1.5 1/310 1.0 x 1.0 1/3

Table 5.2: List of target boiling data runs.

tuations, as opposed to some other unmeasured common-mode noise source, a battery of

runs were taken in which the parameters of the beam and target were varied. Table 5.2

details the conditions for these runs, while Figure 5.14 displays the amount of boiling noise

calculated for each run, using Equation 5.9.

The term “spot size” refers to the dimensions of the beam at the entrance to the scatter-

ing chamber, as determined by the wire array (Section 3.4). Smaller beam spot size should

translate to greater target boiling effects. The vast majority of production data was taken

in the 1.0 mm x 1.0 mm spot size configuration. The pump speed regulates the velocity of

the liquid hydrogen around the target loop. Slower speeds should naturally translate into

larger boiling effects since the hydrogen remains in the beam for longer periods of time.

The nominal pump speed produces a liquid hydrogen velocity of 10 m/s, while the lowest

pump speed produces essentially zero flow [41].

Figure 5.14 shows little variation in the amount of calculated boiling noise among the

configurations tested. The small data spread indicates that the dominant common-mode

noise between the lumi and the Møller detector is not caused by target density fluctuations.

Even if all of the common-mode noise between the two detectors were due to boiling, the

effect is still only roughly 50 ppm. This amount of boiling noise is well below the proposal

Page 120: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

97

Figure 5.14: Extracted boiling noise.

goal of 100 ppm [37]. Barring a large asymmetry observed by the lumi, this level of noise

renders boiling effects inconsequential for the Møller detector analysis.

5.9 Synchrotron Radiation Suppression

As discussed in Section 5.2, it is important that synchrotron radiation comprise no more

than 1% of the total lumi signal. Simulations show that the aluminum material in front of

the lumi reduces the synchrotron background by a large factor. To measure the suppression,

the signal sizes with the target in place and empty target runs are compared. When the

target is out, the dominant background is synchrotron radiation. Figure 5.15 displays the

synchrotron suppression results obtained for each chamber. The channel numbering follows

the convention established in Figure 5.5.

As expected from the synchrotron distribution (Figure 5.4) the background is heavily

peaked in chamber two, the channel on the right side of the lumi rings. Even if all of

this background were due to synchrotron photons, it is clear that it is significantly below

Page 121: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

98

Figure 5.15: Synchrotron radiation background levels by chamber.

the desired 1% level. Because there are assuredly other background sources present, the

true synchrotron radiation background is actually closer to the difference between chamber

two and chamber zero, roughly 0.2%. The maximum asymmetry expected from synchrotron

radiation is at the 600 ppb level, so the asymmetry contribution of the synchrotron radiation

is reduced to the level of a few ppb.

5.10 Linearity

Although the lumi signal flux is quite large, the detector was designed to remain linear to

better than 1%. The main design concern is that electrons and ions could recombine before

being detected on the plates in the chambers. This effect would likely scale with signal size,

producing a non-linearity. The small distance between the collection plates was chosen to

limit electron transit time, and the 100 V bias was chosen to produce a fast signal. These

two design features limit the chance of electron-ion recombination, and are expected to

produce a very linear response.

Page 122: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

99

5.10.1 Requirements

The requirement that individual systematic uncertainties be at the level of 5 ppb or below

can be used to set the linearity limit for the E158 detectors. For simplicity, the response of

the detector to charge is assumed to be a blend of linear and quadratic terms through

S = αF − βF 2, (5.10)

where S is the lumi signal size, F is the real signal flux, and α and β are constants. The

flux is related to the beam charge N through

FL,R = σL,RN, (5.11)

where L and R refer to the beam helicity, and σL,R is proportional to the scattering cross

section. The non-linearity of the detector response ε is defined as the ratio of the quadratic

term to the full signal, as

ε ≡ βF 2

αF − βF 2≈ βF

α. (5.12)

The E158 toroids are known to be very linear (Section 3.3.2), so their response T is assumed

to be simply proportional to charge:

T = N. (5.13)

Inserting the forms in Equations 5.10 through 5.13 into the asymmetry defined in Equa-

tion 5.2 yields

AMeasured = (1 − ε)APhys − εAToroid, (5.14)

Page 123: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

100

where APhys refers to the true parity-violating asymmetry, and AToroid is the charge asym-

metry measured with the toroid. However, because charge is included as a regression pa-

rameter in Equation 5.3, the term proportional to AToroid is effectively removed. Note that

the charge regression coefficient can be identified as ε, the non-linearity of the detector. The

final result is then

AMeasured = (1 − ε)APhys. (5.15)

Assuming that the physics asymmetry is 15 ppb for the lumi and 150 ppb for the Møller

detector, the linearity 1− ε must be determined to 30% for the lumi and 3% for the Møller

detector, to keep the systematic uncertainty at the level of 5 ppb. Perhaps surprisingly, this

renders the linearity of the lumi unimportant.

The charge regression coefficient for the lumi varies between 3% and 8%. At face value,

this indicates that the lumi is less linear than expected. However, while the charge coefficient

certainly contains non-linearity information, as discussed in Section 6.7.2.2, it also contains

sensitivity to any unmeasured beam parameter. Therefore, while regression against charge

removes detector non-linearity, it can not be used to quantify it.

5.10.2 Measured Linearity

As demonstrated in the preceding section, it is not paramount that the lumi possess excep-

tional linearity, because of the additional regression performed against charge. However,

since the detector was designed to keep non-linearities below the level of 1%, it is still

instructive to quantify the linearity of the detector.

In principle, the linearity could be determined by observing the ratio of lumi signal to

the toroid signal for various currents, since the toroids are known to be essentially linear.

Figure 5.16 displays the lumi signal versus current for a single data run.

Page 124: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

101

Figure 5.16: Lumi signal versus beam current.

Because most beam parameters such as position and angle are also correlated to charge,

Figure 5.16 is not a clean representation of the lumi response to current. The lumi is sensitive

to beam motion, so plots such as this can only be used to bound the lumi non-linearity to

∼10%, which is similar to the result obtained from the charge regression coefficient.

To obtain a better linearity determination, the effect of the motion of the beam must be

mitigated. One way to do this is to observe the ratio of the unnormalized lumi asymmetryA0

to toroid asymmetry AToroid. The asymmetries are defined as

ALumi0 ≡ SR − SL

SR + SL(5.16)

and

AToroid ≡ TR − TL

TR + TL, (5.17)

with S and T referring to the lumi and toroid signals, respectively. Using the forms for the

lumi and toroid signals given in Equations 5.10 and 5.13, the relation between ALumi0 and

Page 125: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

102

AToroid is

ALumi0 = (1 − ε)AToroid. (5.18)

The ε term that appears on the right-hand side of Equation 5.18 modifying the proportion-

ality to the toroid asymmetry is again the lumi non-linearity.

The linearity is determined by plotting the ratio of ALumi0 to AToroid as a function of the

analysis cut on beam jitter. As the beam cuts are tightened, the ratio of the uncorrected

lumi asymmetry to the toroid asymmetry then tends toward 1 − ε, the linearity of the

detector.

Figures 5.17 and 5.18 show the results of measuring this ratio as a function of the cut on

beam jitter. Here 100 production data runs have been used, which is roughly one quarter

of the full data set for Run I. The beam cut refers to the maximum allowed beam jitter,

in both X and Y, measured at the lumi. These numbers were obtained using the lumi as a

BPM, as detailed in Section 5.12. The points at the left of the plot have the most stringent

beam jitter cut.

Figure 5.17: Full range of results.

It is clear that the ratio never departs dramatically from unity, indicating that no gross

Page 126: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

103

Figure 5.18: Tight beam cuts subset of linearity results.

non-linearity exists. As the beam cuts are tightened, the ratio is seen to vary between 98.0%

and 100.5%, demonstrating the linearity of the lumi to the level of ≈ 2%.

5.11 Missing Pulse Effect

An unexpected feature of the lumi signal reveals itself after a beam pulse is absent. Missing

pulses occur regularly at 0.5 Hz, due to a scheduled pedestal pulse, and random beam drop-

outs also occur. Figure 5.19 displays the charge-normalized front lumi signal versus time,

with a missed-pulse occurring at the zero of the x-axis. The signal is noticeably larger after

the missed pulse, though it is still only 0.5% above the average signal. The effect decays

quickly and is virtually absent after 3 pulses.

A possible explanation for this behavior is that the detector, which is really a capacitor, is

only allowed to recharge fully during the missing pulse. This would explain why subsequent

pulses receive a boost in size. No other detectors see the effect, so it appears that this

behavior is internal to the lumi design.

Whatever the true cause, the effect produces additional tails to the lumi asymmetry

Page 127: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

104

Figure 5.19: Charge normalized lumi signal following a missing pulse.

distributions, resulting in an effective decrease in the asymmetry resolution. To combat

this, a cut in the analysis is made to remove the first 8 pulses after the pedestal pulse.

This amounts to a 7% cut on the data. Figure 5.20 demonstrates the improved detector

resolution when the additional cut is included.

Figure 5.20: The lumi asymmetry distribution without (left) and with (right) a cut after amissing pulse. RMS is the statistical width of the distribution, while Sigma refers to thewidth determined from the fit.

This cut renders the data sets of the lumi and Møller detector slightly different. So that

the lumi can still serve as a viable systematic cross-check, it was verified that the cut does

Page 128: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

105

not change the mean returned by the Møller detector at a significant level.

5.12 Lumi as a BPM

There are no amplifiers used for the lumi chambers, and all chambers are identical, making

the relative gain among the chambers very close to one. An interesting application of this

feature allows the lumi to be used as a beam position monitor.

The X and Y positions at the lumi are calculated by weighting the chamber signals S

by the chamber spatial positions through

X = αx

∑8n=1 sin θSn∑8n=1 Sn

(5.19)

and

Y = αy

∑8n=1 cos θSn∑8n=1 Sn

, (5.20)

where θ refers to the azimuthal location of the chamber. By convention, zero degrees refers

to the chamber at the top of the lumi ring, and the angle increases in the clockwise direction

looking downstream from the target.

The α coefficients are determined by using the beam position in the BPMs upstream of

the target and extrapolating to the position at the lumi detector. Several runs were taken

with large angular displacement on target during the course of the experiment, allowing for

a large lever-arm to determine the α coefficients. Figure 5.21 displays the results obtained

utilizing these runs. Each point represents the average position over the course of one full

run.

The displacements used in these plots are an order of magnitude greater than would be

Page 129: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

106

Figure 5.21: Calculated beam position using the lumi, versus position position at lumifigured using angle and position bpms.

expected naturally during a typical data run. It is clear that the lumi response to smaller

displacements should be linear. Note that the α coefficients are different for X and Y

because the beam was not centered through the detector.

The main utility of using the lumi as a beam monitor is that it provides the beam

location at the position of the detectors. Having an effective BPM was particularly useful

when adjusting the dipole chicane magnets, which are downstream of the last E158 BPMs.

Also, it was simple to create a display that continually updated the position reported by

the lumi, so that data takers could quickly see that the beam was well positioned through

the detectors. Moreover, since the detector gains were found to be very stable, the lumi

proved to be a useful tool for beam re-alignment after long beam-off periods.

Page 130: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

107

Chapter 6

Asymmetry Analysis

This chapter describes the determination of the parity-violating asymmetry APV from the

raw asymmetry ARaw measured with the Møller detector. The effects of background cor-

rections and dilutions are discussed, and the systematic uncertainty on the measurement is

estimated. The analysis of the luminosity monitor data is also presented.

6.1 Initial Processing

Before any analysis is performed, an initial pass is made over the data to put it into a usable

form. The 0.5 Hz no-beam pulses are used to subtract the pedestal from all detector and

beam monitor channels. The subtraction for a given beam pulse utilizes the running average

of the preceding 10 pedestal pulses. The data from the Møller detector is also “blinded”

with a random offset, such that the asymmetry A from each channel becomes

A→ A+A′, (6.1)

where A′ is a single random number in the range of ±200 ppb. The value of A′ was not

revealed until the asymmetry analysis was completed. The offset was included to reduce

the potential for human bias in the determination of APV .

Page 131: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

108

6.2 Detector Channel Weights

The raw pulse-pair asymmetry ARaw measured with the Møller detector is defined as a

weighted average of the detector channels, given by

ARawi =30∑j=1

wjAji . (6.2)

Here Aji denotes the asymmetry measured with the jth detector channel for the ith pulse

pair, and wj is the weight for that channel. The weights wj are constrained by

30∑j=1

wj = 1. (6.3)

The sum covers the 30 channels of the Møller detector, comprised of the In and Mid regions

(Section 4.1) of the E158 calorimeter depicted in Figure 4.4.

The weights are found on a per-run basis through a minimization method which takes

into account inter-channel correlations. The method is employed to maximize the asym-

metry resolution of the overall Møller detector. First, the symmetric matrix Mij is defined

by

Mij = 〈AiAj〉 − 〈Ai〉〈Aj〉, (6.4)

where An denotes the asymmetry measured with channel n. Channel weights are defined

by minimizing the sum N , defined as

N =30∑i=1

30∑j=1

wiwjMij . (6.5)

Note that in the absence of correlations among the channels in Equation 6.4, the weights

Page 132: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

109

reduce to

wi =1σ2

i∑30j=1

1σ2

j

, (6.6)

the standard statistical weights.

6.3 Regression

The sensitivity of the detectors to beam motion and energy jitter is reduced through a

process called regression. Figure 6.1 depicts the raw asymmetry of the Møller detector

(Equation 6.2) versus the Y position asymmetry, measured with a BPM just upstream of

the target. The strong correlation indicates that asymmetries in the Y position can be

manifested in the asymmetry measured with the Møller detector.

Figure 6.1: Møller detector asymmetry versus Y position asymmetry.

The regression procedure involves calculating the correlation slope seen in Figure 6.1

and removing it on a per-pulse basis. The detector asymmetry ARaw is corrected for the

beam asymmetry ∆Bn through

A = ARaw −6∑i=1

cn∆Bn, (6.7)

Page 133: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

110

where A is the corrected asymmetry, and the sum runs over the beam parameters of charge

Q, energy E, X and Y position, and X and Y angle. The regression coefficients cn are the

correlations of the detector asymmetry with the nth beam parameter.

The coefficients are experimentally determined by performing a multidimensional fit of

ARaw against the six beam parameters. The covariance matrix Bij is defined as

Bij = 〈∆Bi∆Bj〉 − 〈∆Bi〉〈∆Bj〉, (6.8)

and the vector Vi is

Vi = 〈Araw∆Bi〉 − 〈ARaw〉〈∆Bi〉. (6.9)

The coefficients cn are obtained by inverting the matrix B and multiplying by the vector

V .

cj =6∑i=1

B−1ij Vi (6.10)

Because the coefficients could change over time, they are determined every 10,000 pulse

pairs.

Figure 6.2 depicts the same data presented in Figure 6.1 after the regression correction

of Equation 6.7 has been applied. The correlation of the detector asymmetry with the beam

asymmetry is greatly reduced.

Regression also enhances detector resolution by reducing the effect of beam fluctuations

on the asymmetry distribution measured with the detector. Figure 6.3 depicts the im-

provement in detector performance with regression. The resolution is improved by ≈ 60%.

The different E158 detectors are positioned at different geometries, causing their sensi-

Page 134: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

111

Figure 6.2: Møller detector asymmetry versus Y position asymmetry after regression.

Figure 6.3: Møller detector resolution with and without regression.

tivity to beam parameters to vary greatly. Table 6.1 presents the coefficients of three E158

detectors, averaged over all Run I data. The Out detector is the most sensitive detector

for all categories. The large coefficients are due to the large slope in its signal flux profile

(Figure 3.25), so that small deviations in beam parameters result in comparatively large

changes in the Out detector signal size.

Page 135: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

112

Run 1 DetectorCoefficient Møller Out Front Lumi

Q (ppm/ppm) 0.005 -0.08 -0.03E (ppm/MeV) -25.9 76.2 -5.1X (ppm/µm) 0.5 6.8 -1.1Y (ppm/µm) -1.2 2.0 0.4dX (ppm/µR) -66.1 128.2 25.2dY (ppm/µR) 7.3 65.5 -12.9

Table 6.1: Regression coefficients for Run I.

6.4 Beam Dithering

Beam dithering is an alternative method to regression for correcting detector asymmetries.

The coefficients of Equation 6.7 are found by intentionally moving the beam with corrector

magnets and observing the detector correlation with the beam monitors. Figure 6.4 depicts

Figure 6.4: Location of components used for beam dithering.

the location of the linac components used for beam dithering. Corrector magnets located in

the final sectors of the linac are used to vary both beam position and angle, by an amount

several times the natural beam jitter. The energy is varied by adjusting the phase of a

klystron in the same region. Figure 6.5 depicts the response of one of the lumi chambers to

an x-position dithering cycle1.

The beam parameters were typically dithered 4% of the time. This data is only used

for the determination of correction slopes and is removed from the production data. The1The slowly varying dither cycle depicted in the figure was found to be very disruptive to the beam

feedbacks. For the production data, a more rapid dithering style was employed.

Page 136: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

113

Figure 6.5: A lumi chamber responding to a position dither cycle.

response of the detectors to the dithered beam parameters is found in a manner similar to

the regression method, by a multidimensional fit to the data in which dithering is occurring

(Equation 6.10). The detector asymmetry is then corrected as in Equation 6.7. Because

both the lumi and Out detector require a correction for charge, a regression against charge

is done after the other beam coefficients are computed.

In contrast to the regression procedure, dithering varies the beam parameters in a charge

independent way. By decoupling the parameters, the dithering results are less prone to

systematic effects. However, dithering was not always functional, while the regression pro-

cedure can always be performed. Therefore, the main results for the E158 experiment are

computed using regression, while the dithering results are used as a cross-check.

6.5 Data Selection

The quality of the data used in the asymmetry analysis is ensured by removing pulses that

do not pass predetermined criteria [39]. To avoid biasing the data, many of the data cuts

also remove a certain number of pulses before and after an offending pulse. The following

paragraphs describe each of the requirements employed in the asymmetry analysis.

Page 137: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

114

General: The primary quality cut is composed of several separate cuts that are always

made in the analysis. The first ensures that the beam parameters of charge, energy, X

and Y position, and X and Y angle are within six standard deviations of the mean value.

The second removes all data with large excursions due to beam dithering. The third cut

ensures that the time between beam pulses is the same for each pulse in the pulse pair, to

combat possible hysteresis effects. The final cut requires that the agreement of the charge

asymmetry measurements of the toroids upstream of the target be within 1000 ppm, roughly

20 standard deviations. Any pulse pair not meeting these criteria is removed, as well as the

50 preceding and subsequent pulse pairs. Overall, this cut retains 90.8% of the data.

Position Jitter: Pulse pairs with position jitter greater than nine standard deviations

from zero are removed, as well as the 50 preceding and following pulse pairs. The cut has

a 91.9% retention rate.

Energy Stability: The momentary loss of a klystron in the linac is not uncommon. De-

pending on the phase of the missing klystron, the energy of the beam can either increase or

decrease, usually by an amount of ≈ 200 MeV. Off-energy pulses are removed by checking

for large deviations in the E158 energy BPM. The cut has a window of ± 50 pulse-pairs,

and has a retention rate of 97.7%.

BPM Phase: The resolution of the BPMs decreases when the phase of the local oscillator

drifts. This cut uses the Q cavity of each BPM to ensure that the phase is within an

acceptable range. The time structure for the phase drifts causes this cut have a window

from 5 pulse pairs before the offending pulse to 50 pairs after. The retention rate is 97.4%.

Page 138: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

115

BPM Linearity: Large offsets, on the scale of a few millimeters, cause the BPM response

to become non-linear. This cut ensures that the beam position is within a range where the

BPMs are linear to better than 99%. The cut window is from two pulse pairs before the

offending pulse to 4 pulse pairs after. The retention rate is 97.0%.

Charge: The beam charge is required to be greater than 1011 electrons per pulse, about

20% of the normal beam current. Pulses with low charge are fundamentally different from

the standard beam pulses, and are excluded. The cut only removes one pulse pair. The

retention rate is 98.1%.

Transmission: The transmission from the source to the E158 target is required to be

greater than 90%. Like low-current pulses, low-transmission pulses are very different from

normal beam pulses and are removed. The retention rate is 99.9%.

Source Voltages: Data in which the CP and PS Pockels cell voltages are improperly set

at the source are excluded by this cut. When the voltages are incorrect, the polarization of

the beam is unknown. The retention rate is 99.3%.

Timeslot: While the beam repetition rate is 120 Hz, the electric power for the accelerator

is at the normal 60 Hz rate. The position of the beam pulses in time relative to the phase

of the electric power can therefore assume two states, which are called timeslots. Because it

is possible that the two timeslots have different properties, this cut requires that each pulse

in a pair occur in the same timeslot. The retention rate is 98.9%.

Slopes: The regression coefficients used to correct the detector asymmetries are calculated

in chunks for 10,000 pulse pairs, sorted by timeslot. It is possible that one timeslot can

dominate a period of time so that the other timeslot has very few pulses. If the smaller

Page 139: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

116

timeslot has less than 100 pairs, this cut removes that timeslot because its regression slopes

would be unreliable. The retention rate is 99.2%.

All Cuts The total data retention rate when all cuts are employed is 73.5%.

Careful studies were done on the Run I Møller detector asymmetry result in which cuts

were included or removed to observe their effect on the mean asymmetry. In all cases it was

found that the cuts do not bias the data in any discernible way.

6.6 Møller Detector Asymmetry Analysis

The regression corrected Møller detector asymmetry distribution defined by Equation 6.7

is used to determine a mean asymmetry Ai and statistical uncertainty σi for each data run

consisting of roughly 200,000 pulse pairs. The results of N individual data runs can be

combined to compute the average asymmetry Aaverage and uncertainty σaverage, with

Aaverage =

∑Ni=1

Ai

σ2i∑N

i=11σ2

i

(6.11)

and

σaverage =

√√√√ 1∑Ni=1

1σ2

i

. (6.12)

Figures 6.6 and 6.7 present the results obtained with the Møller detector for all of Run

I and Run II, respectively. The data has been averaged into periods of data called “slugs,”

which are runs with the same beam energy and source halfwave plate setting. (The different

energy and halfwave plate states are used to combat systematic effects, and are described

in the following section.) Both Run I and Run II are comprised of approximately the same

amount of data, though Run II has less systematic reversals.

Page 140: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

117

Figure 6.6: Run I Møller detector asymmetry versus slug.

The Run I and Run II plots have reasonable χ2/ndf and their means agree within

statistical fluctuations. The clear asymmetry in both plots represents the first observation

of parity violation in Møller scattering.

The data plotted in Figures 6.6 and 6.7 are purely statistical averages that do not include

systematic uncertainties such as beam polarization, regression correction uncertainties, and

background corrections. These issues are discussed in Section 6.8 to arrive at APV .

6.6.1 Systematic Reversals

The experiment utilized two distinct methods to reduce sensitivity to helicity-correlated

systematic effects [37]. The first involved running with two different source halfwave plate

states (Section 3.1.4). Inserting the waveplate reverses the helicity of the laser light hitting

the cathode while keeping the rest of the source setup the same. The second method

involved running at two separate beam energies, 45 GeV and 48.3 GeV. The difference in

these energies represents a 180 g-2 rotation of the electron spin as it traverses the A-Line

Page 141: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

118

Figure 6.7: Run II Møller detector asymmetry versus slug.

bend leading up to End Station A.

The four possible apparatus configurations combine the physics asymmetry APV with

two types of helicity-correlated systematic effects denoted as A1sys and A2

sys. The asymmetry

A1 refers to systematic effects that are not affected by the waveplate state, such as electronics

cross-talk with the bias voltage of the CP cell at the source (Section 3.1). The asymmetry

A2 refers to effects like residual linear polarization in the laser light at the source that can

also reverse sign with the insertion of the source halfwave plate. The four configurations

are then given by

A45Out = APV +A1

Sys +A2Sys

A45In = APV −A1

Sys +A2Sys

A48Out = APV −A1

Sys −A2Sys (6.13)

A48In = APV +A1

Sys −A2Sys,

Page 142: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

119

(6.14)

where AEnergyHalfwave denotes the overall asymmetry seen with the Møller detector for each

energy and halfwave plate setting. By collecting the same amount of data in each state,

systematic effects are minimized. Run II had less systematic reversals than Run I and will

be shown to have larger systematic effects, discussed in Section 6.7.2.2.

Figure 6.8 presents the Møller detector asymmetry obtained for each energy-halfwave

plate configuration in Run I and Run II. All of the states agree well, indicating that there

are no large systematic effects. Figures 6.9 and 6.10 present the same data as the previous

plots, but without the systematic reversal sign flips taken into account. They provide

another visual indication that systematic uncertainties are below the level of statistical

fluctuations for the Møller detector.

Figure 6.8: Measured asymmetry for each energy-halfwave plate setting.

6.6.2 Beam Corrections

Because beam monitors are used to modify the raw Møller detector asymmetry through

Equation 6.7, beam asymmetries lead to a correction of the detector asymmetry. Table 6.2

Page 143: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

120

Figure 6.9: Run I Møller detector asymmetry versus slug, sign flips suppressed.

Run IParameter Correction (ppb)

Charge 0.2Energy 5.1

X 6.7Y 0.05

X Angle 23.8Y Angle 2.0Total 37.9

Run IIParameter Correction (ppb)

Charge -1.8Energy 29.0

X -3.7Y -10.5

X Angle 17.1Y Angle -8.9Total 21.2

Table 6.2: Beam corrections to ARaw of the Møller detector.

lists the beam corrections for Run I and Run II. Note that the corrections have no un-

certainty: they are exactly the amount the beam monitors have shifted the raw Møller

detector asymmetry. The contribution to the uncertainty on the Møller asymmetry enters

through the systematic uncertainty associated with the method itself and is discussed in

Section 6.7.2.2.

The agreement between the corrected detector asymmetry obtained with regression and

dithering is an important measure of potential systematic uncertainties. Table 6.3 com-

pares the asymmetries obtained with the two methods. The values differ slightly from the

Page 144: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

121

Figure 6.10: Run II Møller detector asymmetry versus slug, sign flips suppressed.

Run I Asymmetry (ppb) Correction (ppb)Regression -179.1 ± 24.3 30.6Dithering -182.2 ± 27.0 27.5Run II Asymmetry (ppb) Correction (ppb)

Regression -150.0 ± 22.3 11.5Dithering -154.8 ± 22.7 6.7

Table 6.3: Comparison of regression and dithering results for the Møller detector asymmetry.

averages shown in Figure 6.6 and 6.7 because only the subset of data in which dithering

was functional was used. The improvement in the Møller detector resolution is less for

dithering, as indicated by the uncertainty on the asymmetry. The most likely cause is that

the dithering correlation slopes are computed less often than the regression slopes. In both

Run I and Run II, the methods agree to the level of a few ppb, indicating that systematic

differences between the methods are small. However, Section 6.7.2.2 will demonstrate that

the total systematic uncertainty contribution to APV due to beam corrections is dominated

by a systematic effect common to both regression and dithering.

Page 145: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

122

6.7 Asymmetry Corrections and Uncertainties

The parity-violating asymmetry APV is obtained from the beam corrected Møller detector

asymmetry AMeasured by accounting for contributions from backgrounds and taking into

account scale factors. The asymmetries are related through

APV =1Pλ

AMeasured −∑i fiAi1 −∑i fi

, (6.15)

where the sums are taken over all backgrounds. The fi terms represent the dilution factors of

the backgrounds, defined as the ratio of the background signal to the total signal seen in the

Møller detector. The helicity-dependent asymmetries of the backgrounds are represented

by the Ai terms. The scale factors are the beam polarization P and the linearity of the

detector λ. Sections 6.7.1 through 6.7.8 detail the components of Equation 6.15 for the

Møller detector. Section 6.8 then uses the results to compute APV for Run I and Run II.

6.7.1 The Electron-Proton Scattering Correction

The spectrometer (Section 3.7) is designed to provide separation of electron-electron and

electron-proton (eP) scattering events. However, the separation is not perfect, and some

electron-proton events are registered in the Møller detector. In particular, inelastic eP

events are expected to have an asymmetry an order of magnitude larger than APV . It is

therefore crucial to account for the eP background contamination in AMeasured.

The amount of eP signal in the Møller detector is determined from a Monte Carlo sim-

ulation of the E158 spectrometer [43]. The simulation is configured to match the maps of

the signal flux versus radius, provided by the profile monitor (Section 4.3). Figure 6.11

depicts a comparison between the simulation and a profile scan with typical running condi-

Page 146: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

123

tions. Figure 6.12 presents a scan with the insertable acceptance collimator (Section 3.7.4)

in place. The additional separation of the Møller from the electron-proton scattering events

is important for ensuring that the simulation matches the reality of the signal distribution.

In both plots, the agreement between the simulation and the actual data is reasonable.

Figure 6.11: Data and simulation comparison with normal running conditions.

Because the difference between the data and the simulation is small, the simulation can

be used to quantify the eP contamination in the Møller detector. Figure 6.13 displays a

profile scan with the electron-electron and electron-proton contributions separated in the

simulation.

Table 6.4 presents the dilution factors f used for determining the correction to AMeasured

for the electron-proton contribution in Equation 6.15. The ratio of elastic to inelastic events

R is also shown. The uncertainty on the results was determined by varying the input

parameters of the simulation within a reasonable range.

The Run I and Run II results differ for two reasons. First, the eP collimator (Sec-

Page 147: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

124

Figure 6.12: Data and simulation comparison with the insertable acceptance collimator inposition.

tion 3.7.5) was only used in Run II. The collimator blocks the eP detector and most of

the Out ring. Second, the quadrupole magnet settings of the spectrometer were changed

between Run I and Run II to de-emphasize the Out ring of the detector.

The value of AMeasured must also be corrected for the asymmetry contribution of the

electron-proton scattering background. The asymmetry result of the eP detector in Run I

is used as an input to the simulation. The flux in the eP detector is approximately 70%

Run I 45 GeV 48 GeVDetector f σf R f σf R

In 0.0969 0.0099 7.4 0.0861 0.0087 6.8Mid 0.0684 0.0079 5.7 0.0610 0.0071 5.1Out 0.1401 0.0203 4.4 0.1903 0.0261 4.0

Run II 45 GeV 48 GeVDetector f σf R f σf R

In 0.0810 0.0095 7.1 0.0780 0.0079 6.1Mid 0.0540 0.0067 5.8 0.0520 0.0053 5.5Out 0.1020 0.0143 4.9 0.0810 0.0153 4.4

Table 6.4: Dilution factors f due to background eP scatters. R is the ratio of elastic toinelastic eP signals.

Page 148: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

125

Figure 6.13: Profile scan with the Møller and electron-proton scattering simulation resultssuperimposed.

Run I 45 GeV 48 GeVDetector fA σfA fA σfA

In -33.4 4.7 -35.5 4.8Mid -32.2 5.2 -33.8 5.8Out -86.9 22.3 -140.2 33.1

Run II 45 GeV 48 GeVDetector fA σfA fA σfA

In -29.4 4.6 -33.2 4.6Mid -26.5 3.8 -27.5 3.8Out -57.5 9.0 -52.4 9.2

Table 6.5: Diluted asymmetries fA due to the electron-proton scattering background. Allentries are in ppb.

elastic eP scatters, 27% inelastic eP scatters, and only about 3% Møller scattered electrons.

However, the asymmetry associated with the inelastic eP events is roughly 25 times larger

than the elastic asymmetry, so these events dominate.

Figure 6.14 depicts the asymmetry measured with the eP detector for Run I. The Q2 is

sufficiently different between the two beam energies that they are averaged separately. The

result is consistent with the theoretical estimate [68].

Table 6.5 details the diluted asymmetry fA due to the electron-proton scattering back-

ground used to correct AMeasured. The large uncertainty on the Run I result for the Out

ring is dominated by the simulation of the shower sharing with the eP detector. In Run II,

the eP collimator greatly suppresses the effect, reducing the systematic uncertainty.

Page 149: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

126

Figure 6.14: Asymmetry result from the eP detector in Run I.

6.7.2 Beam Asymmetry Correction Systematic Uncertainties

The regression corrections given in Equation 6.7 must be assigned a systematic uncertainty

contribution to APV . Examining Equation 6.15, the uncertainty can be accommodated

formally as a correction with zero dilution f or asymmetry fA, but with an uncertainty

σfA.

The systematic uncertainty estimate is divided into two pieces: first-order and higher-

order effects. The following sections will demonstrate that while the In and Mid regions of

the detector are fairly insensitive to the beam correction systematic effects, the Out ring

is dominated by the higher-order effects. Because of the large uncertainty from this effect,

the Out ring is not included with the In and Mid ring in the calculation of APV .

6.7.2.1 First-Order Beam Correction Systematic Uncertainties

The first-order systematic uncertainty assigned to the values in Table 6.2 is found by en-

hancing the Møller detector sensitivity to the regression parameters by re-weighting the

Page 150: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

127

Regression CoefficientsParameter Monopole X-Dipole Y-Dipole

Q (ppm/ppm) 0.005 0.012 0.005E (ppm/MeV) -25.9 -2.6 -10.0X (ppm/µm) 0.5 -22.1 -0.8Y (ppm/µm) -1.2 -1.6 24.0dX (ppm/µR) -66.1 -30.8 -31.0dY (ppm/µR) 7.3 18.9 151.7

Table 6.6: Regression coefficients of three Mid detector weighting schemes.

detector channels. In general, the asymmetry A for any weighting pattern is defined as

A =1√

N∑Ni=1w

2i

N∑1

wnAn, (6.16)

where An refers to the asymmetry of channel n, with weight wn [70]. The sums are taken

over the N channels of the detector. The most relevant weighting scheme is the dipole

pattern, with weights given by

wDipolen = sin(2πnN

). (6.17)

Assigning n = 0 as the top of the detector, this scheme produces enhanced sensitivity

to beam motion in the horizontal direction and is known as the X-Dipole pattern. Rotating

by 90, the pattern becomes a Y-Dipole. Table 6.6 displays the regression coefficients of

the Mid ring Monopole, Mid X-Dipole, and Mid Y-Dipole weighting schemes.

The monopole weighting scheme is dominated by the energy parameter, while the X

and Y dipoles are dominated by the spatial parameters their names suggest. Owing to their

increased sensitivity, specific weighting schemes are employed to estimate the systematic

uncertainty inherent in the regression method. Table 6.7 relates the patterns that were

especially sensitive to each of the beam parameters2.2The Q parameter is excluded in the first-order systematic uncertainty estimate. It is more relevant for

the higher-order uncertainty discussed in Section 6.7.2.2.

Page 151: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

128

Sensitive Weighting SchemesParameter Detector

E Møller MonopoleX Mid X-DipoleY Mid Y-DipoledX Out X-DipoledY Out Y-Dipole

Table 6.7: Sensitive pattern-weighted detectors for the regression beam parameters.

The first order systematic uncertainty on the beam corrections is estimated by using

the sensitive monitors in Table 6.7 to compute the relative error εn on each of the n beam

corrections. Two separate sets of data, denoted by the superscripts, are used to determine

the relative uncertainty through

εn =A1n −A2

n

corr1n − corr2n→ ∂An

∂corrn, (6.18)

where An is the mean asymmetry of the detector weighting pattern sensitive to the nth

beam parameter, and corrn refers to the beam correction of An. To ensure that first order

effects dominate, the two data sets are chosen as the two timeslots in the data slug with

the largest difference in the beam corrections.

The first-order systematic uncertainty δn for the correction corrn is the product of the

relative uncertainty with the value of the correction:

δn = εncorrn. (6.19)

Tables 6.8 and 6.8 display the systematic uncertainty calculated for each regression

parameter using this method for both Run I and Run II. To be conservative, the value

quoted as the relative error is the magnitude of the error found using Equation 6.18, plus

one standard deviation. The uncertainty on the total correction is found as the quadrature

Page 152: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

129

Run IParameter Correction (ppb) Relative Error ε (%) Uncertainty δ (ppb)

E 5.1 7.0 0.36X 6.7 8.7 0.58Y 0.1 10.1 0.01dX 23.8 3.9 0.92dY 2.0 22.3 0.45

Total: 1.2

Table 6.8: Run I first-order systematic uncertainties in the regression corrections to the rawMøller detector asymmetry.

Run IIParameter Correction (ppb) Relative Error ε (%) Uncertainty δ (ppb)

E 29.0 10.0 2.00X -3.7 3.8 0.17Y -10.5 2.8 0.31dX 17.1 3.6 0.69dY -8.9 5.7 0.55

Total: 3.0

Table 6.9: Run II first-order systematic uncertainties in the regression corrections to theraw Møller detector asymmetry.

sum of the individual uncertainties. In both data sets, the total first-order systematic

uncertainty is small.

6.7.2.2 Higher-Order Beam Correction Systematic Uncertainties

The plots of the Møller asymmetry versus slug shown in Figures 6.6 and 6.7 have χ2/ndf

near unity, and the systematic reversal plots of Figure 6.8 show no indication that large

systematic uncertainties are present. However, when the Out ring is examined in the same

manner, it is clear that there are possible systematic effects that need to be addressed.

Figures 6.15 and 6.16 present the Out ring asymmetry data. The Run I slug plot

has an elevated χ2/ndf, but still looks reasonable. The Run I systematic reversal plot

appears to have some cancellation of an effect between the 48 GeV-Halfwave Out and 45

GeV-Halfwave Out detector results. In the Run II plots, however, it is clear that there

Page 153: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

130

are systematic effects not removed by the regression method. The degree to which this

unregressed systematic effect can influence the Møller detector asymmetry must be assigned

a systematic uncertainty.

Figure 6.15: Run I Out ring asymmetry data.

Figure 6.16: Run II Out ring asymmetry data.

The nature of the systematic effect was isolated in Run III of the E158 experiment. In

addition to the normal setup, the BPM signals were also “sliced” in time and fed to four

different ADC channels. That is, the first quarter of the pulse went to the first channel,

and so on. Effectively, this allows for regression against the shape of the beam pulse. The

Run III Out ring data look similar to Run II, with poor fits when the normal analysis is

employed. However, when regression against the sliced BPM signals is included, the Out

Page 154: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

131

data are greatly improved, with a χ2/ndf near unity [72]. Therefore, the cause of the Out

detector systematic effect is found to be intra-pulse helicity-correlated asymmetries which

are not removed by the normal regression procedure in Run I and Run II.

The nature of the effect can be understood by the following simple description. The

BPMs measure the average position X of the electron beam pulse3, given by

X =1T

∫ T

0x(t)dt, (6.20)

where the integral is over the duration of the pulse T and x is the beam position. The

regression procedure then corrects the raw detector asymmetry ARaw by multiplying the

beam difference ∆X by a constant coefficient c, as in Equation 6.7. The correction C is

then given as

C =c

T

∫ T

0x1(t) − x2(t)dt, (6.21)

where the subscripts refer to the beam pulses of the pulse pair. The method is valid unless

the detector sensitivity to the beam varies over the course of the beam pulse. In that case,

the constant c becomes a function of position, and actually needs to be integrated as

C =1T

∫ T

0c(x1)x1(t) − c(x2)x2(t)dt. (6.22)

Effectively, this is what “slicing” the BPM signals does: it allows for the detector sensitivity

to vary over the course of the beam pulse.

A possible mechanism for this to occur is known as “tail wagging,” where the tail of

the beam pulse has large fluctuations due to the passage of the head of the pulse through3Technically, the BPMs measure the charge-averaged position. The formula could be corrected by x(t)

→ ρ(t)x(t), where ρ is the charge density. However, for clarity, the charge dependence is omitted.

Page 155: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

132

the linac. This is known as the “wakefield effect,” where the electromagnetic “wake” of the

head of the pulse disrupts the tail [73, 74]. The effect could be on the scale of hundreds

of microns, which would indeed produce large effects in the Out ring, due to its sensitive

geometry [75]. If the intra-pulse effects are also helicity-correlated, the asymmetry results

are not fully corrected, as seen in the Out ring results in Figure 6.15 and 6.16.

Because there were no “sliced” BPM signals in Run I and Run II, the systematic un-

certainty due to higher-order effects can only be estimated. Because the plots of the Møller

detector asymmetry versus slug look normal, it is expected that the sensitivity to higher-

order effects should be small. First, it is assumed that the measured asymmetry AMeasured

is related to the real asymmetry AReal through

AMeasured = AReal + αAsys, (6.23)

where Asys is the asymmetry due to beam-related systematic effects, and α is the detector

sensitivity to Asys [76]. If two detectors are examined, Equation 6.23 can be used to find

α1Asys ≡ σ1

sys =α1

α1 − α2((AMeasured

1 −AMeasured2 ) − (AReal1 −AReal2 )), (6.24)

where the subscripts denote the detectors and σ1sys is the systematic uncertainty assigned

to detector 1. If we choose the Møller detector as the first detector and the Out ring as the

second, we have

AReal1 −AReal2 = ∆eP, (6.25)

where ∆eP represents the difference in the correction for the electron-proton scattering

background between the two detectors. The only unknown values are then α1 and α2,

Page 156: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

133

Systematic Uncertainty Estimates (ppb)Run I Run II

∆eP 60.6 ± 26.1 21.1 ± 5.6AMeasured

1 −AMeasured2 -69.7 ± 46.5 -104.1 ± 48.5

α1α2

-0.064 -0.058σMøllersys -7.8±3.2 -6.9±2.7

Table 6.10: The higher-order systematic uncertainty computed for the Møller detector,comparing the Møller detector (1) and the Out ring (2), assuming Equation 6.26.

which can only be estimated.

One method is to note that the Out ring has a large charge regression coefficient (Ta-

ble 6.1), between five and ten times larger than would be expected (Section 5.10.1). Because

most beam parameters are correlated to charge, it is reasonable to assume that the charge

coefficient is a measure of the size of the systematic effect.

As an estimate, it is assumed that the α systematic coefficient is proportional to the

charge regression coefficient cQ. The ratio of the sensitivity to the systematic effect of the

two detectors is then

α1

α2=c1Qc2Q. (6.26)

With Equation 6.25 and the assumption of Equation 6.26, the systematic uncertainty con-

tribution for the Møller detector can be found with Equation 6.24. Table 6.10 represents

the Møller systematic uncertainty found for Run I and Run II using this method.

Because Equation 6.26 is an assumption, it is useful to estimate the ratio α1α2

by an

alternative method. The uncertainty on AMeasured in Equation 6.23 is

σ2AMeasured = σ2

AReal + α2σ2Asys. (6.27)

Page 157: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

134

Systematic Uncertainty Estimates (ppb)Run I Run II

∆eP 60.6 ± 26.1 21.1 ± 5.6AMeasured

1 −AMeasured2 -69.7 ± 46.5 -104.1 ± 48.5

α1α2

0 ± 0.062 -0.141 ± 0.016σMøllersys 0 ± 8.1 -15.5 ± 6.2

Table 6.11: The higher-order systematic uncertainty computed for the Møller detector,comparing the Møller detector (1) and the Out ring (2), using Equation 6.29.

Again using two detectors denoted by 1 and 2, the ratio of the α coefficients is

∣∣∣∣α1

α2

∣∣∣∣ =√√√√√σ2

AMeasured1

− σ2AReal

1

σ2AMeasured

2− σ2

AReal2

, (6.28)

which can be approximated as

∣∣∣∣α1

α2

∣∣∣∣ = σAMeasured1

σAMeasured2

√√√√χ2ndf1 − 1χ2ndf2 − 1

. (6.29)

Geometry causes the regular regression coefficients for the Møller and Out detectors

to be anti-correlated, so it is reasonable to assume that α1α2

is negative. When one of the

detectors has a χ2ndf < 1, the method breaks down, because no systematic effects can be

discerned. Therefore, it is also useful to calculate the uncertainty on Equation 6.29:

σ

(α1

α2

)=σAMeasured

1

σAMeasured2

√1

2ndfχ2ndf1 − 1χ2ndf2 − 1

√1

(χ2ndf1 − 1)2

+1

(χ2ndf2 − 1)2

. (6.30)

Using Equations 6.29 and 6.30, a table similar to Table 6.10 can be constructed. Table 6.11

presents the results found with this method.

To be consistent with the two methods, the systematic uncertainty estimate for Run I

is set to 10 ppb, while for Run II it is 15 ppb. The data in Run III is still being analyzed,

but preliminary findings indicate that the results of Table 6.10 and Table 6.11 are an

Page 158: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

135

overestimate of the higher-order beam correction systematic uncertainty for the Møller

detector [72].

6.7.3 Dipole Asymmetry

A transversely polarized electron beam scattering off unpolarized electrons produces an

asymmetry with azimuthal dependence at the Møller detector. The effect is purely electro-

magnetic, due to a two photon exchange box diagram [77]. Data taken at 46 GeV, where

the beam was measured to be 85% transversely polarized in the horizontal direction, clearly

show the effect. Figure 6.17 plots the asymmetry measured in the Mid ring versus channel

number for the transverse polarization data.

Figure 6.17: Run I asymmetry plotted versus channel number for 46 GeV running.

Production data is taken with a longitudinally polarized beam, though residual trans-

verse polarization at the level of a few percent is possible. Figure 6.18 depicts the asymmetry

versus azimuth taken from the production data. Comparing the amplitudes of Figures 6.17

and 6.18 indicates that the beam is ≈ 5% transversly polarized during normal production

running, consistent with reasonable expectations.

The effect is purely electromagnetic, and so averages exactly to zero over the full az-

imuth. However, the Møller channels are statistically weighted, spoiling the azimuthal

symmetry of the detector. The dipole pattern of Figure 6.18 can then contaminate the

Page 159: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

136

Figure 6.18: Run I asymmetry plotted versus channel number for production data.

Møller asymmetry result. Because the signal flux distribution is well centered on the face

of the Møller detector, the degree to which the dipole pattern shifts the asymmetry result

of each ring of the detector δ can be quantified by

δX = AX

N∑i=1

wi sin(2πiN

) (6.31)

and

δY = AY

N∑i=1

wi cos(2πiN

), (6.32)

where A refers to either the X or Y amplitude of the oscillation in Figure 6.18, the weights

w are defined by Equation 6.2, and the sum is taken over the N channels in the ring.

The oscillation in Figure 6.18 is dominated by the Y-dipole, due to the precession of the

longitudinal polarization in the horizontal bends of the A-Line. However, the A-Line also

has several vertical bends, incorporating a slight X-dipole oscillation into the observed

distribution. Table 6.12 gives the shift to the asymmetry measured with the Møller detector

Page 160: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

137

Dipole Shifts (ppb)45 GeV 48 GeV

Run I -11.7 ± 3.4 -2.9 ± 1.9Run II -11.3 ± 3.4 0.4 ± 2.3

Table 6.12: The computed shifts in AMeasured due to the dipole amplitude from transversebeam polarization.

due to the dipole oscillation, computed for both Run I and Run II.

6.7.4 Spot Size Sensitivity

The detector sensitivity to beam spot size is quantified by measuring the correlation with

the wire array, as described in Section 3.4. For both Run I and Run II, the systematic

shift to AMeasured due to the spot size asymmetry was found to be consistent with zero.

A systematic uncertainty of 1 ppb is assigned to the possible systematic effect, with no

correction made.

6.7.5 Pion Background

The pion detector (Section 4.4) is used to estimate the asymmetry and dilution factor due

to pions interacting in the Møller detector. The Run I result was found to be contaminated

by particles passing through the eP detector portion of the E158 calorimeter. In Run II, the

insertable eP collimator blocked this background, making the pion detector result viable.

The Run II result will be used for both Run I and Run II. The measured dilution factor f

is 0.0014 ± 0.0011 ppb, with an asymmetry correction fA of -0.5 ± 0.8 ppb.

6.7.6 Synchrotron Radiation

The synchrotron light monitor (SLM) is used to determine the vertical polarization of the

electron beam (Section 4.5). The vertical polarization, in turn, induces a helicity-dependent

Page 161: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

138

Diluted Synchrotron Asymmetry (ppb)45 GeV 48 GeV

Run I -1.6 ± 1.2 1.5 ± 1.2Run II -1.6 ± 1.2 2.8 ± 2.2

Table 6.13: Synchrotron correction calculated with the SLM.

asymmetry in the synchrotron radiation produced in the E158 spectrometer chicane. The

radiation produces a dilution factor f of 0.0015 ± 0.0005 for both Run I and Run II. The

radiation also represents an asymmetry correction, shown in Table 6.13.

An alternative analysis computes the vertical polarization of the beam through observing

the amplitude of the X-dipole oscillation in the plot of asymmetry versus azimuth in the

Møller detector. The oscillation is due to the two photon exchange box diagram, discussed

in Section 6.7.3. The method has the advantage that it computes the vertical polarization at

the target, whereas the SLM is located in the middle of the A-Line bend. This analysis finds

that the asymmetry contribution fA in Run I is 0.0 ± 4.5 ppb, and 0.0 ± 2.3 ppb in Run

II. The results are consistent with the SLM analysis, and they are used in the calculation

of APV .

6.7.7 Neutral Backgrounds

Neutral hadrons produced in the E158 calorimeter can penetrate the lead shielding of the

PMTs in the detector and produce a signal by directly impacting the photocathodes. Con-

tamination from hadrons produced in the eP detector that make it into the Møller PMTs

must be quantified, due to their large asymmetry.

Data was taken with PMTs blinded with aluminum tape, so that they were only sen-

sitive to neutral backgrounds. By comparing data with the quadrupole magnets off and

with the insertable eP collimator in place or removed, it was determined that the neu-

tral hadron background from the eP detector represents a dilution factor of 0.003 ± 0.001

Page 162: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

139

for the Møller detector. The asymmetry correction is just the dilution multiplied by the

asymmetry measured in the eP detector, yielding -5 ppb ± 3 ppb, with the uncertainty

set conservatively [78]. The background is only important for Run I, because it is greatly

suppressed in Run II by the eP collimator.

A second neutral background of concern is high-energy photons that bounce from spec-

trometer collimators into the Møller detector. Data taken with the spectrometer quadrupole

magnets off determine that this background represents a dilution factor of 0.004 ± 0.002 [79].

The quadrupoles-off runs also found a large asymmetry in the Møller detector of 2.5

ppm. The most likely cause is electron-Aluminum scattering in the beam pipe. Simulations

indicate that the effect is reduced by a factor of three when the quadrupole magnets are

on. Assuming a worst-case scenario, where the entire dilution factor of 0.004 is due to this

type of background, the asymmetry correction is 2.5 ppm × 0.3 × 0.004 = 3 ppb. Due to

the large uncertainty in this analysis, a 100% uncertainty is applied to the correction.

6.7.8 Scale Factors

The beam polarization directly scales AMeasured into APV . The polarization was determined

with the polarimeter (Section 4.2.4) to be 84.9% ± 4.6% in Run I and 84.4% ± 4.6% in

Run II.

Section 5.10.1 demonstrates that the detector linearity will also directly scale the mea-

sured value of APV . The linearity of the Møller detector is found by weighting the linearity

of the In and Mid rings (Section 4.1.4) by their statistical power, yielding an overall linearity

of 99.0% ± 1.1%, in both runs.

Page 163: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

140

Run IDilutions and Corrections

45 GeV 48 GeVfA σfA f σf fA σfA f σf

Beam 1st Order 0.0 1.2 0.0000 0.0000 0.0 1.2 0.0000 0.0000Beam 2nd Order 0.0 10.0 0.0000 0.0000 0.0 10.0 0.0000 0.0000

Dipole Bias -11.7 3.4 0.0000 0.0000 -2.9 1.9 0.0000 0.0000eP Correction -32.7 5.0 0.0795 0.0087 -34.5 5.4 0.0707 0.0077

Spot size 0.0 1.0 0.0000 0.0000 0.0 1.0 0.0000 0.0000Synchrotron 0.0 4.5 0.0015 0.0005 0.0 4.5 0.0015 0.0005

High-Energy Photons 3.0 3.0 0.0040 0.0020 3.0 3.0 0.0040 0.0020Neutral eP Leakage -5.0 3.0 0.0030 0.0010 -5.0 3.0 0.0030 0.0010

Pions -0.5 0.8 0.0014 0.0011 -0.5 0.8 0.0014 0.0011Total -46.9 13.3 0.0894 0.0090 -39.9 13.2 0.0806 0.0081

Scale FactorsBeam Polarization 0.849 ± 0.046Detector Linearity 0.990 ± 0.011

Table 6.14: The dilution factors f are dimensionless, while the asymmetry corrections fAare given in units of ppb.

6.8 Determination of APV

The results presented in Sections 6.7.1 to 6.7.8 are used with Equation 6.15 to obtain APV

from AMeasured. Tables 6.14 and 6.15 contain all of the dilution factors and asymmetry

corrections applied in Run I and Run II respectively. Table 6.16 presents the experimen-

tally measured value of APV at a Q2 of 0.026 (GeV/c)2, the primary result of the E158

experiment. The final chapter converts APV into sin2θW for comparison to theory, and uses

it to set limits on new physics phenomena.

6.9 Luminosity Monitor Results

The data from the luminosity monitor (Chapter 5) is analyzed in much the same way as the

Møller detector [81]. The expected asymmetry in the lumi signal flux is -15 ppb ± 5 ppb, an

order of magnitude smaller than APV . The detector serves as a sensitive null-asymmetry

cross-check for the results of the Møller detector. The data presented in this section is

Page 164: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

141

Run IIDilutions and Corrections

45 GeV 48 GeVfA σfA f σf fA σfA f σf

Beam 1st Order 0.0 3.0 0.0000 0.0000 0.0 2.2 0.0000 0.0000Beam 2nd Order 0.0 15.0 0.0000 0.0000 0.0 15.0 0.0000 0.0000

Dipole Bias -11.3 3.4 0.0000 0.0000 0.4 2.3 0.0000 0.0000eP Correction -27.7 4.1 0.0648 0.0078 -29.7 4.1 0.0707 0.0077

Spot size 0.0 1.0 0.0000 0.0000 0.0 1.0 0.0000 0.0000Synchrotron 0.0 2.3 0.0015 0.0005 0.0 2.3 0.0015 0.0005

High-Energy Photons 3.0 3.0 0.0040 0.0020 3.0 3.0 0.0040 0.0020Neutral eP Leakage 0.0 0.0 0.0000 0.0000 0.0 0.0 0.0030 0.0010

Pions -0.5 0.8 0.0014 0.0011 -0.5 0.8 0.0014 0.0011Total -36.5 16.7 0.0717 0.0082 -26.8 16.5 0.0691 0.0067

Scale FactorsBeam Polarization 0.844 ± 0.046Detector Linearity 0.990 ± 0.011

Table 6.15: The dilution factors f are dimensionless, while the asymmetry corrections fAare given in units of ppb.

APV (ppb)Run I 45 GeV -183 ± 38(stat) ± 20(sys)Run I 48 GeV -158 ± 48(stat) ± 19(sys)

Run I Combined -173 ± 30(stat) ± 20(sys)

Run II 45 GeV -159 ± 44(stat) ± 23(sys)Run II 48 GeV -131 ± 37(stat) ± 22(sys)

Run II Combined -143 ± 28(stat) ± 23(sys)

TOTAL -158 ± 21 (stat) ± 17 (sys)

Table 6.16: The parity-violating asymmetry in Møller scattering.

Page 165: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

142

Run 1 DetectorCoefficient Channel 00 Channel 02 Full Lumi

Q (ppm/ppm) -0.55 -0.80 -0.03E (ppm/MeV) -20.9 -95.7 -5.1X (ppm/µm) -7.9 55.3 -1.1Y (ppm/µm) -25.2 -4.1 0.4dX (ppm/µR) -389.8 1459.5 25.2dY (ppm/µR) -392.8 697.6 -12.9

Table 6.17: Comparison of the regression coefficients for two individual lumi channels andthe full detector. Channel 00 is at the top of the lumi ring, while channel 02 is on the rightside.

for the front lumi ring only, for simplicity. The back ring has poorer resolution, and is

consistent with the front ring in all cases.

The lumi channels are identical ion chambers, which are automatically gain matched.

The sensitivity of the detector to beam parameters is greatly suppressed when the channels

are added with equal weight, in contrast to the Møller detector in Equation 6.2. The lumi

asymmetry ARaw is the average of the individual channel asymmetries Ai, given by

ARaw =18

8∑i=1

Ai. (6.33)

The sensitivity suppression is clear in Table 6.17, which compares the regression coefficients

for two individual lumi chambers with the coefficients for the whole detector.

Table 6.1 illustrates that the lumi has sensitivity more like the Møller detector than

the Out ring, except for the conspicuous exception of the Q coefficient. The inflated Q

coefficient again indicates that not all beam parameters are being completely regressed for

the lumi. Figures 6.19 and 6.20 plot the lumi asymmetry data for Run I and Run II. The

χ2/ndf is elevated in Run I and very large in Run II, similar to the behavior of the Out

ring.

Because the lumi asymmetry is expected to be essentially zero, the mean of the system-

Page 166: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

143

Figure 6.19: Run I lumi asymmetry data.

Figure 6.20: Run II lumi asymmetry data.

atic reversal plots, with the sign flips suppressed, is indicative of the size of the underlying

systematic effect (Figure 6.21). The Run II reversal plot indicates that the unregressed

systematic effect is at the level of 70 ppb. The effect is much smaller in Run I, most likely

due to the higher frequency of systematic reversals in the Run I data set.

6.9.1 Lumi Beam Correction Systematic Uncertainty

It is clear from the systematic reversal plots that higher-order beam correction systematic

uncertainties are potentially important for the luminosity monitor, though systematic re-

versals appear to have provided a large suppression factor. Unfortunately, because the lumi

and Out ring both have large Q coefficients, the systematic uncertainty estimation methods

Page 167: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

144

Figure 6.21: Run I and Run II lumi asymmetry with systematic reversal sign flips ignored.

outlined for the Møller detector in Section 6.7.2.2 do not have enough statistical power to

set useful limits.

The systematic uncertainty estimate for the beam corrections to the lumi is found

through a less direct manner. First, “composite detectors” are defined by combining weight-

ing schemes of various detectors Ai to produce a composite detector Cn that is sensitive to

beam parameter n, while insensitive to all others, through

Cn =N∑i=1

wiAi, (6.34)

where the sum is taken over the N detectors used, and the weights wi are chosen to en-

hance the sensitivity of the composite monitor to a single beam parameter [80]. Table 6.18

presents a simple example of the enhanced sensitivities of a composite monitor4: the Out-

Mid combination has an enhanced sensitivity to energy fluctuations compared to the Out

or Mid detector alone.

More complicated composite monitors can be constructed to precisely isolate the sen-4The angle parameters have been converted from nRadians to nm to make their relative size with the

other beam parameters more clear.

Page 168: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

145

Coefficient ComparisonParameter Mid Out Out - Mid

E (ppm/MeV) -26.6 76.2 102.8X (ppm/µm) 0.5 6.8 6.3Y (ppm/µm) -0.1 2.0 2.1dX (ppm/µm) -1.7 3.2 4.9dY (ppm/µm) 0.6 1.6 2.2

Table 6.18: The composite monitor Out-Mid has enhanced energy sensitivity.

Composite Monitor Coefficientsppb/keV ppb/nm

Composite E X Y dX dYCE 102.3 -0.1 -0.1 0.0 0.0CX -0.1 -22.0 0.0 -0.6 0.4CY 0.0 -0.1 -22.6 0.0 0.0CdX -0.1 0.0 0.6 24.1 0.0CdY -0.1 -0.1 0.7 0.0 -34.2

Table 6.19: Regression coefficients of the composite monitors Cn.

sitivity to one beam parameter. The composite monitors are made up of combinations of

the front lumi, the In, Mid, and Out rings of the Møller detector, and the X and Y dipole

weighting scheme for each detector. It is expected that the unregressed systematic effects

will be greatly enhanced in the composite monitors.

Table 6.19 details the composite sensitivities achieved with this method. The composite

monitors Cn are each composed of roughly seven of the original detectors. The In, Mid,

and Out ring are always added in a way to ensure that the physics asymmetry APV cancels

between them, leaving only the difference in the eP corrections.

The asymmetry measured with each composite monitor is calculated and compared to

the expected value. The expectation can be non-zero, due to differences in the eP corrections

of the Møller detector regions, and the vertical polarization of the beam can induce real

dipole asymmetries. The systematic uncertainty δn assigned to the regression correction of

Page 169: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

146

Run I Front Lumi Systematic Uncertainty EstimatesParameter ∆ cComp clumi suppression α systematic error (ppb)

E 92.3 ± 50.1 102.3 -5.1 -0.050 -4.6 ± 2.5X -55.8 ± 26.5 -22.0 -1.1 0.050 -2.8 ± 1.3Y -139.2 ± 53.9 -22.6 0.4 -0.018 2.5 ± 1.0dX 18.2 ± 60.2 24.1 0.6 0.025 0.5 ± 1.5dY 117.5 ± 70.9 -34.2 -0.3 0.009 1.0 ± 0.6

TOTAL: -3.4 ± 6.9

Table 6.20: Run I systematic uncertainty estimates for the front lumi. The columns labeledcComp and clumi contain the dominant regression coefficient for the composite monitor andthe lumi, respectively.

Run II Front Lumi Systematic Uncertainty EstimatesParameter ∆ cComp clumi suppression systematic error (ppb)

E 120.8 ± 44.6 67.4 -5.8 -0.086 -10.4 ± 3.8X 63.8 ± 21.0 -14.6 -1.4 0.096 6.1 ± 2.0Y -46.7 ± 41.6 -22.4 -0.8 0.036 -1.7 ± 1.5dX 123.0 ± 60.2 25.2 2.9 0.115 14.2 ± 6.9dY 135.2 ± 75.8 -33.4 0.6 -0.018 -2.4 ± 1.4

TOTAL: 5.8 ± 15.6

Table 6.21: Run II systematic uncertainty estimates for the front lumi. The columns labeledcComp and clumi contain the dominant regression coefficient for the composite monitor andthe lumi, respectively.

the nth parameter for the lumi is then

δn = αn(CMeasuredn − CExpectedn ) → αn∆n. (6.35)

The coefficient αn is the ratio of the the nth lumi regression coefficient to the sensitive

coefficient of the composite monitor Cn. The ratio expresses the relative sensitivity of the

lumi monitor to the composite monitor for beam parameter n. Table 6.20 and Table 6.21

relate the systematic uncertainties calculated for the lumi in this manner for Run I and

Run II. The total systematic uncertainty is found by combining the individual systematic

uncertainties and errors linearly, to be conservative.

The uncertainty assigned for beam corrections for the front lumi is 7 ppb in Run I and

Page 170: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

147

16 ppb in Run II. The uncertainty is still quite small compared to the variations observed

in Figure 6.21. It is clear that the energy and halfwave plate systematic reversals provided

large cancellation of the unregressed systematic effect and were critical for the lumi result.

6.9.2 Lumi Dipole Contamination

Unlike the Møller detector, the lumi channels are not weighted in forming the asymmetry of

the whole detector, so the dipole contamination shifts in Equation 6.31 and Equation 6.32

from the Møller detector analysis in Section 6.7.3 are identically zero. However, since the

lumi is not necessarily aligned with the center of the Møller detector, and hence the beam

path, a large position offset at the lumi could still induce a dipole shift. Figure 6.22 displays

the individual channel asymmetries for the lumi in Run I. There is no indication of any large

offset, since the sinusoid fits well. Therefore, no systematic uncertainty will be assigned to

the lumi for this effect.

Figure 6.22: Run I lumi asymmetry plotted versus channel number.

Page 171: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

148

6.9.3 Lumi Dilutions and Scale Factors

The measured lumi asymmetry must be corrected for backgrounds and uncertainties through

Equation 6.15. The only background for the lumi comes from synchrotron radiation. In

Section 5.9, the synchrotron dilution factor f , averaging over the whole detector, is found

to be 0.0035 ± 0.0004. The uncertainty is set by the target-out signal seen in the vertical

chambers.

The asymmetry correction for the synchrotron background would require a detailed

simulation of the lumi to determine its analyzing power for vertical polarization. This

analysis is underway, but is not yet completed. However, the result is expected to be very

similar to the Møller detector result, so the Møller detector value is taken as an estimate,

scaled by the ratio of the dilution factors for the two detectors. In Run I, the synchrotron

asymmetry correction is 0.0 ± 10.5 ppb, and in Run II it is 0.0 ± 5.4 ppb.

The only additional asymmetry correction is due to sensitivity to the beam spot size.

The spot size systematic was found to be ±1 ppb for Run I and ±2 ppb for Run II in

Section 3.4.

The lumi has the same two scale factors as the Møller detector: beam polarization and

detector linearity. The linearity used is 99 ± 2%, based on the results of Section 5.10.

6.10 Lumi Asymmetry Result

Combining the corrections and systematic uncertainties of Sections 6.9.1 and 6.9.3 with

the measured asymmetries in Section 6.9 yields the final luminosity monitor asymmetry.

Table 6.22 presents the lumi results. Both the Run I and Run II data are consistent with

the predicted value of -15 ppb ± 5 ppb. The result is a consistency cross-check for the

Page 172: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

149

Luminosity Monitor Asymmetry ResultsRun I -19.6 ppb ± 17.6 ppb (stat) ± 15.2 ppb (sys)Run II -16.5 ppb ± 14.1 ppb (stat) ± 20.4 ppb (sys)Total -18.1 ppb ± 11.0 ppb (stat) ± 17.6 ppb (sys)

Table 6.22: Luminosity monitor asymmetry results.

asymmetry measured with the Møller detector.

Page 173: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

150

Chapter 7

The Weak Mixing Angle

This chapter presents the result for sin2θW calculated from the SLAC E158 measurement

of the parity-violating asymmetry APV in Møller scattering. The result is then used to set

limits on possible extensions to the Standard Model. Finally, future experiments planned

to measure the running of sin2θW are discussed.

7.1 Extraction of sin2θW

The QED calculation of APV is covered in Chapter 2. Using the convention for sin2θeffW

given in Equation 2.14, the asymmetry is given by

APV =−GµQ2

√2πα

1 − y

1 + y4 + (1 − y)4Fb(y)(1 − 4 sin2 θeffW ), (7.1)

where y is related to the center-of momentum scattering angle θCM through

y =1 − cos θCM

2. (7.2)

The scale factor Fb(y) is included to account for bremsstrahlung effects that depend

on the geometry of the experiment [82]. The relevant diagrams are depicted in Figure 7.1.

Page 174: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

151

Bremsstrahlung radiation before scattering effectively lowers the beam energy, while radiat-

ing afterward affects the experimental acceptance. The overall effect of Fb was determined

to be small, with an average value of 1.01 ± 0.01.

Figure 7.1: Bremsstrahlung diagrams included in Fb(y). The crossed versions must also becomputed, for a total of 16 diagrams.

The analyzing power AP is defined as the flux weighted average over the kinematic

factors in Equation 7.1 so that

APV = AP(1 − 4 sin2 θeffW ). (7.3)

The analyzing power is computed using the same Monte-Carlo simulation that was employed

for background subtraction in Section 6.7.1. Table 7.1 presents the computed values for

AP, where In and Mid refer to particular regions of the Møller detector (Section 4.1). The

uncertainty on AP was found by varying the parameters of the simulation within reasonable

bounds. The simulation was also used to determine that the average Q2 is 0.026 (GeV/c)2.

Page 175: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

152

Analyzing Power (ppm)Run I 45 GeV 48 GeV

In -2.976 ± 0.060 -3.171 ± 0.060Mid -3.304 ± 0.053 -3.539 ± 0.042

Run II 45 GeV 48 GeVIn -3.046 ± 0.061 -3.182 ± 0.060

Mid -3.372 ± 0.044 -3.537 ± 0.042Average: -3.298 ± 0.051

Table 7.1: Analyzing powers computed from simulation. The overall average AP is deter-mined by weighting the entries with the corresponding uncertainty on APV .

Combining the average analyzing power from Table 7.1,

−3.298 ppm ± 0.051 ppm, (7.4)

and the measured value of APV from Table 6.16,

−158 ppb ± 21 ppb (stat) ± 17 ppb (sys), (7.5)

the weak mixing angle is found to be

sin2 θeffW = 0.2380 ± 0.0016 (stat) ± 0.0013 (sys). (7.6)

For comparison, the theoretical prediction is

sin2 θeffW = 0.2385 ± 0.0006 (theory). (7.7)

It is clear that the agreement between the experimentally measured value and the theoretical

Page 176: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

153

prediction is quite good. The difference between them is

∆ sin2 θeffW = −0.0005 ± 0.0021, (7.8)

corresponding to ≈ 0.25 standard deviations.

Equivalently, using the convention for sin2θW (Q2) preferred by theorists (Equation 2.13),

the E158 result is

sin2 θMSW = 0.2376 ± 0.0020. (7.9)

Figure 7.2 presents the E158 value for sin2θMSW on a plot along with the theoretically pre-

dicted Q2 dependence. Figure 7.3 presents the measurements of the weak mixing angle

evolved to the Z0 mass, assuming the Standard Model, to aid in the comparison.

Figure 7.2: Experimental results and the theoretical running of the weak mixing angle.

Page 177: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

154

Figure 7.3: The measured weak mixing angle evolved to the Z0 mass.

7.2 Physics beyond the Standard Model

The deviation of the measured value of sin2θeffW from the theoretical predication can be

used to set limits on possible extensions of the Standard Model. The following sections use

the deviation in Equation 7.8 to set the E158 constraints on new physics.

7.2.1 Electron Compositeness Limit

In the current model of particle physics, the electron is assumed to be a truly point-like

particle. However, it is possible that the electron has substructure [83]. In electron-electron

scattering, electron compositeness can be cast as a contact interaction, with the general

Lagrangian given as

L =4π2Λ2

[ηLLψLγµψLψLγµψL + ηRRψRγµψRψRγ

µψR + 2ηRLψRγµψRψLγµψL]. (7.10)

The Λ term parameterizes the energy scale at which the contact interaction becomes im-

portant. The indices L and R refer to the chiral components of the interaction, signifying

Page 178: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

155

left or right helicity. The η coefficients are free parameters, with the constraint that the

largest member have a magnitude of unity or less. There are no symmetry requirements

that the interaction should conserve parity, so the η values are independent.

The Lagrangian produces an amplitude that is purely real, and at low Q2 there is an

interference term with the dominant electromagnetic diagrams. In contrast, at the Z0

resonance where the dominant amplitude is purely imaginary, no interference term results.

Low energy experiments are therefore inherently more sensitive to new physics in the form

of Equation 7.10 than the Z0 pole experiments.

The process described by Equation 7.10 produces a shift in the predicted value of sin2θW

given by

∆ sin2 θeffW =π√2GF

ηLL − ηRRΛ2

. (7.11)

For parity-violation experiments, it is conventional to set ηLL to ±1 and ηRR to zero to

obtain limits on Λ±LL. The E158 compositeness constraints are then

Λ+LL > 8.0 TeV/c2 (7.12)

and

Λ−LL > 6.9 TeV/c2, (7.13)

at the 95% confidence level.

The present limit on this quantity is from the combined data of the Aleph, Delphi, L3,

and Opal detectors at LEP [84]. They report a constraint on compositeness of

Λ+LL > 8.3 TeV/c2 (7.14)

Page 179: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

156

and

Λ−LL > 10.3 TeV/c2. (7.15)

Individual results from these experiments yield constraints in the range of three to five

TeV/c2. The E158 limit is quite competitive with the collider results.

7.2.2 Scalar Doubly Charged Higgs Limit

Møller scattering is uniquely sensitive to s-channel scattering through exchange of an exotic

doubly charged scalar Higgs boson. Particles such as this often appear in theories beyond

the Standard Model and are not required to respect parity [85]. Figure 7.4 depicts the

diagram for this process.

Figure 7.4: Doubly charged Higgs particle exchange diagram.

The mass of the Higgs particle is constrained to be much larger than the interaction

energy, so it can be described as a contact interaction in the same formalism as the electron

compositeness limit. The process would shift sin2 θeffW by

∆ sin2 θeffW =√

28

(g

MH

)2 ηLL − ηRRGF

, (7.16)

where g/MH is the ratio of the coupling to the mass of the Higgs particle. Again, ηLL is

set to ±1 while ηRR = 0 to obtain the limit on(

gMH

)2

LL± . The 95% confidence limits from

Page 180: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

157

the E158 result are (g

MH

)2

LL+< 0.017GF (7.17)

and (g

MH

)2

LL−< 0.022GF . (7.18)

The current limits are set by observing the related process of muonium to anti-muonium

conversion [86]. Figure 7.5 depicts the diagram for this process. These experiments place

Figure 7.5: Muonium to anti-muonium conversion.

the constraint (g

MH

)2

< 0.14GF , (7.19)

on the doubly charged Higgs parameters, at the 90% confidence level.

The E158 limit represents an order of magnitude improvement. However, it should be

noted that the muonium experiments would still be sensitive even in the absence of parity

violation.

7.2.3 Extra Z Boson Limit

Many extensions to the Standard Model include new heavy analogues to the Z0. The new

particle is usually denoted as a Z′. Some Grand Unified Theories embed the Standard

Model in the group E6, producing two new Z′ states [87]. The extra Z particles assume the

Page 181: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

158

forms [88]

Zβ = Zχ cos β + Zψ sin β (7.20)

and

Zβ′ = −Zχ sin β + Zψ cos β. (7.21)

The Zχ state has distinct couplings to different fermions, while the Zψ state couples uni-

versally to all fermions [89]. The presence of these new particles scales the prediction for

APV [32] by

1 + 7

m

2Z0

m2Zβ

(cos2 β +

√53

sinβ cos β

)+m2Z0

m2Zβ′

(sin2 β −

√53

sin β cos β

) . (7.22)

The mixing angle β is model dependent. In the SO(10) model, a group of considerable

theoretical interest, β = 0 and Equation 7.22 simplifies to

1 + 7m2Z0

m2Zχ

. (7.23)

The weak mixing angle would be modified as

1 − 4 sin2 θeffW →(

1 + 7M2Z0

M2Zχ

)(1 − 4 sin2 θeffW

). (7.24)

The E158 measurement of the weak mixing angle can then be used to set the limit on the

mass of the extra Z particle of

MZχ > 410 GeV/c2, (7.25)

at the 95% confidence level. The single highest current constraint is provided by CDF which

reports Zχ > 595 GeV/c2 at the 95% confidence level [90].

Page 182: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

159

7.2.4 Oblique Parameter X Limit

Extensions to the Standard Model could be manifested through radiative corrections to

the gauge boson propagators, known as oblique corrections. The standard Peskin-Takeuchi

parameters STU were introduced to quantify deviations from the Standard Model due to

oblique corrections for energies much greater than MZ0 [91]. This formalism was extended

to include new physics effects at the weak scale by the introduction of three new parameters

VWX [92].

The running of sin2θW from the Z0 resonance down to low Q2 can be shown to isolate

the X parameter [93]. The κ(0) term in Equation 2.8 is modified as [32]

κ(0) → (1 − 0.032X)κ(0). (7.26)

The shift in sin2 θeffW is then

∆ sin2 θeffW = −ρκ(0)(0.032X) sin2 θMSW (M2

Z0). (7.27)

The quantities ρ and κ are defined in the theoretical calculation of APV in Sections 2.4.4

and Section 2.4.1. Using the E158 result, it is found that the limit on X is

X = 0.07 ± 0.28. (7.28)

The value of X based on a global STUVWX fit to world data in 1994 has an uncertainty

of 0.58 [94]. The inclusion of the E158 result, and to an equal degree the atomic parity

violation result, will produce a substantial improvement.

Page 183: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

160

7.3 Future Experiments

Several upcoming experiments will be conducted to map out the running of the weak mixing

angle. Figure 7.6 depicts the proposed experiments and the expected final uncertainties.

The following sections give brief descriptions of these experiments.

Figure 7.6: Present and future experiments used to map the running of the weak mixingangle.

7.3.1 E158 Run III

The final run of the E158 experiment, Run III, took place in the Summer of 2003. The

amount of data collected was approximately equal to the sum of Run I and Run II. The

addition of the Run III data is therefore expected to reduce the uncertainty on sin2θW by

a factor of ≈ √2. Additionally, the analysis of the “sliced” BPM signals (Section 6.7.2.2)

is expected to reduce the systematic uncertainty on the Run I and Run II results. Overall,

the E158 result should yield an uncertainty of approximately

δ sin2 θeffW = ±0.0014. (7.29)

Page 184: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

161

7.3.2 QWeak

The QWeak experiment has been approved to run in Hall C at Jefferson Lab, beginning in

2007 [95]. It will operate at a Q2 of 0.03 (GeV/c)2, almost identical to the E158 exper-

iment. QWeak will employ a 1.2 GeV longitudinally-polarized electron beam on a 35 cm

liquid hydrogen target to observe the parity-violating asymmetry in elastic electron-proton

scattering.

In the limit of a small scattering angle, the asymmetry APVQWeak is given by

APVQWeak =−GF

4πα√

2

[Q2(1 − 4 sin2 θW ) +Q4B(Q2)

]≈ −300 ppb. (7.30)

The term B(Q2) is due to the electromagnetic and weak form factors of the proton, and

contributes roughly 1/3 of the total asymmetry. It will be determined from the results of

other experiments, including HAPPEX and SAMPLE.

The goal is to perform a 4% measurement of APVQWeak, which would correspond to a 0.3%

measurement of sin2θW . The dominant sources of systematic uncertainty on APVQWeak are

anticipated to be a 2% uncertainty coming from the determination of B(Q2) and a 1.5%

uncertainty from the beam polarization measurement.

7.3.3 DIS-Parity at JLab

The DIS-Parity experiment has been proposed to run in Hall A at Jefferson Lab [96]. The

experiment will use a longitudinally-polarized electron beam on a 15 cm liquid deuterium

target, to measure the parity-violating asymmetry ADIS in deep-inelastic scattering. The

Q2 will be 2.0 (GeV/c)2, placing it between the NuTeV and the E158 results. The parity-

Page 185: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

162

violating asymmetry is given by

ADIS = (109 ppm)Q2[(−3

2+

103

sin2 θW

)+ Y Rv

(−32

+ 6 sin2 θW

)]∼ −100 ppm,

(7.31)

where Y ≈ 1 contains kinematics information and Rv ≈ 1 depends on quark distributions.

The goal of the first run of the experiment will be a determination of sin2θW to ≈ 1%

using a 6 GeV beam. If a large deviation from the Standard Model prediction is observed,

a second run is planned that would employ the 12 GeV upgraded beam at JLab and run at

a larger scattering angle [97].

7.3.4 The Large Hadron Collider

The Large Hadron Collider (LHC) at CERN is expected to become operational in 2008 [98].

It is designed to produce proton-proton collisions with an energy of 14 TeV. By observing a

large number of Z0 decays, it is expected that sin2θW (M2Z0) will be measured to a precision

of ± 0.0001. Also, the large energy scale will allow for direct searches of extra Z bosons,

superseding all indirect constraints. In particular, the Zχ in the SO(10) model will be

constrained at the level of 4.5 TeV, an order of magnitude improvement.

7.3.5 Next Linear Collider

The high-energy lepton-lepton collisions of the proposed “Next Linear Collider” (NLC)

would be required to measure the weak mixing angle for Q2 much larger than M2Z0 [99]. If

the machine is ever constructed, it is expected that collision energies in the range of 1 to

1.5 TeV will be achieved, as shown in Figure 7.6 [100].

Page 186: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

163

7.4 Conclusion

The E158 result represents the first measurement of parity violation in Møller scattering.

The observed value of sin2 θeffW at a Q2 of 0.026 (GeV/c)2 is in agreement with the Standard

Model prediction for the running of the weak mixing angle. The success of the E158 experi-

ment demonstrates the technical feasibility of measuring small parity-violating asymmetries,

setting the stage for future challenging parity violation experiments that will further probe

the Standard Model of particle physics.

Page 187: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

164

Bibliography

[1] H. Becquerel, p. 52 of Nobel Lectures, Physics 1901-1921, (Elsevier Publishing Company,

Amsterdam, 1967).

[2] J. Chadwick, Verh. d. D. Phys. Ges. 16, 383 (1914).

[3] W. Pauli, “Letter to Tubingen Conference,” Pauli Archive at CERN (1930).

[4] E. Fermi, Nuovo Cim. 11, 1 (1934).

[5] T. D. Lee et al., Phys. Rev. 75, 905 (1949).

[6] T. D. Lee and C. N. Yang, Phys. Rev. 104, 254 (1956).

[7] C. S. Wu et al., Phys. Rev. 105, 1413 (1957).

[8] S. L. Glashow, Nucl. Phys. 22, 579 (1961).

[9] S. Weinberg, Phys. Rev. Lett. 19, 1264 (1967).

[10] A. Salam, p. 367 of Elementary Particle Theory, (Almquist and Wiksells, Stockholm,

1969).

[11] F. J. Hasert et al., Phys. Lett. B46, 121 (1973).

[12] C. Y. Prescott et al., Phys. Lett. B77, 347 (1978).

[13] C. Y. Prescott et al., Phys. Lett. B84, 524 (1979).

Page 188: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

165

[14] G. Arnison et al., Phys. Lett. B122, 103 (1983).

[15] K. Hagiwara et al., Phys. Rev. D66, 010001 (2002).

[16] G. P. Zeller et al., Phys. Rev. Lett. 88, 091802 (2002).

[17] S. C. Bennett and C. E. Wieman, Phys. Rev. Lett. 82, 2484 (1999).

[18] E. A. Paschos and L. Wolfenstein, Phys. Rev. D7, 91 (1973).

[19] G. P. Zeller et al., Phys. Rev. D65, 111103 (2002).

[20] K.P.O. Diener et al., hep-ph/0310364v1 (2003).

[21] S. A. Blundell et al., Phys. Rev. D45, 1602 (1991).

[22] W. J. Marciano and J. L. Rosner, Phys. Rev. Lett. 65, 2963 (1990).

[23] S. J. Pollack and M. C. Welliver, Phys. Lett. B464, 177 (1999).

[24] A. Derevianko, Phys. Rev. Lett. 85, 1618 (2000).

[25] V. A. Dzuba et al., Phys. Rev. A63, 044103 (2001).

[26] A. Derevianko, Phys. Rev. A65, 012106 (2001).

[27] M. G. Koslov et al., Phys. Rev. Lett. 86, 3260 (2001).

[28] V. A. Dzuba et al., Phys. Rev. D66, 076013 (2002).

[29] M. Y. Kuchiev and V. V. Flambaum, hep-ph/0305053 (2003).

[30] E. Derman and W. J. Marciano, Annals Phys. 121, 147 (1979).

[31] T. Kinoshita and A. Sirlin, Phys. Rev 113, 1652 (1959).

Page 189: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

166

[32] A. Czarnecki and W. J. Marciano, Phys. Rev. D53, 1066 (1995).

[33] W. J. Marciano and A. Sirlin, Phys. Rev. D22, 2695 (1980).

[34] T. B. Humensky et al., “Setup of Longpulse Laser System for E158,” E158 Source

Document (2001).

[35] T. B. Humensky et al., Nucl. Inst. Meth. A521, 261 (2004).

[36] T. Maruyama et al., Nucl. Inst. Meth. A492, 199 (2002).

[37] R. Carr et al., SLAC-Proposal-E-158 (1997)

[38] D. Whittum and Yu. Kolomensky, Rev. Sci. Instrum. 70, 2300 (1999).

[39] P. Mastromarino, “Beam Asymmetry Analysis for E-158 Run 1,” E158 Technote 34

(2003).

[40] M. Woodley and M. Woods, “Skew Quad SQ27.5,” E158 Technote 25 (2001).

[41] J. G. Weisend et al., “The Cryogenic System for the SLAC E158 Experiment,” E158

Technote 11 (2001).

[42] J. Gao et al., Nucl. Inst. Meth. A498, 90 (2003).

[43] C. G. Arroyo et al., “eP Inelastic Scattering Asymmetry Correction,” E158 Technote

33 (2003).

[44] M. Woods et al., “Synchrotron Background Analysis for Run I,” E158 Technote 39

(2003).

[45] A. E. Bondar and E. L. Saldin, Nucl. Inst. Meth. 195, 577 (1982).

[46] I. Younus, “E158 Calorimeter,” E158 Technote 47 (2003).

Page 190: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

167

[47] D. Relyea, “Møller Electronics Specifications,” E158 Technote 26 (2001).

[48] M. Cooke, “Probing Møller Detector Linearity with the Polarized Iron Foil and Light

Filters,” E158 Technote 48 (2003).

[49] H. L. B. Gould and D. H. Wenny, Elec. Eng. (Amer. Inst. Elec. Eng.) 76, 208 (1967).

[50] C. Møller, Ann. Phys. (Leipzig) 14, 532 (1932).

[51] N. M. Shumeiko and J. G. Suarez, J. Phys. G26, 113 (2000).

[52] E. Chudakov, “E158 Polarimetry, New Detector,” E158 Technote (2003).

[53] E. Chudakov, “Polarimetry, Run I,” E158 Technote 43 (2002).

[54] L. G. Levchuk, Nucl. Inst. Meth. A345, 496 (1994).

[55] C. Arroyo, “Profile Detector Analysis of E158 Run,” E158 Technote 19 (2001).

[56] D. L’huillier, “Pion Detector Analysis,” E158 Technote 42 (2003).

[57] D. L’huillier, GEANT3 simulation for E158 (unpublished) (2002).

[58] B. T. Tonguc and M. Woods, “Report on Synchrotron Radiation Analysis,” E158

Technote 17 (2001).

[59] S. A. Belomesthnykh, Nucl. Inst. Meth. 227, 173 (1984).

[60] B. Tweedie, “The GEANT3 Simulation of E158,” E158 Technote 37 (2003).

[61] L. Keller, “LUMI Asymmetry and Acceptance as a Function of Transverse and Angular

Displacement at the Target,” E158 Technote 28 (2001).

[62] C. Arroyo, “Data/Simulation Comparison of Cherenkov Detector Scans for Run I,”

E158 Technote 41 (2003).

Page 191: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

168

[63] D. Relyea, GEANT Simulation for E158 (unpublished) (2001).

[64] C. Field, EGS4 Simulation for E158 (unpublished) (2001).

[65] J. D. Jackson, Rev. Mod. Physics 48, 417 (1976).

[66] I. P. Karabekov and R. Rossmanith (Jefferson Lab), CEBAF-PR-93-014 (1993).

[67] A. V. Zarubin, Nucl. Inst. Meth. A283, 409 (1900).

[68] P. Bosted, “Radiative Corrections for E158,” E158 Technote 23 (2003).

[69] D. Relyea, “Run I Asymmetry Analysis,” E158 Technote 53 (2003).

[70] G. M. Jones, “E158 Run I Luminosity Monitor Asymmetry Analysis,” E158 Technote

38 (2003).

[71] K. Bega, “Run II: Summary of Møller Unblinded Results,” E158 Technote 52 (2002).

[72] A. Vacheret, “Run III Higher Orders Systematic Studies,” E158 Technote 59 (2003).

[73] J. Farrell, “Characterization of the E158 Beam,” E158 Technote 57 (2003).

[74] J. T. Seeman, “Linear Collider Accelerator Physics Issues Regarding Alignment,” p.

257 of Stanford 1989 Accelerator Alignment Proceedings, (1989).

[75] P. Bosted, “Predicted Dithering Response for E158,” E158 Technote 49 (2003).

[76] Y. Kolomensky, “Higher Order Asymmetry Systematics,” E158 Technote 56 (2003).

[77] L. Dixon and M. Schreiber, hep-ph/0402221 (2004).

[78] L. Kaufman and K. Kumar, “Neutral Backgrounds in the Møller Detector,” E158

Technote 44 (2003).

Page 192: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

169

[79] K. Kumar, “Update on Neutral Backgrounds,” E158 Technote 55 (2003).

[80] M. Woods et al., “Run I Systematic Error for Beam Asymmetry Corrections,” E158

Technote 45 (2003).

[81] G. M. Jones, “Lumi Asymmetry Analysis, All Data Runs,” E158 Technote 54 (2003).

[82] V. A. Zykunov, Yad. Phys. 66 (2003).

[83] E. J. Eichten et al., Phys. Rev. Lett 50, 811 (1983).

[84] K. Hagiwara et al., Phys. Rev. D66, 010001 (2002).

[85] N. G. Deshpande et al., Phys. Rev. D44, 837 (1991).

[86] V. A. Gordeev et al., Physics of Atomic Nuclei 60, 1164 (1997).

[87] D. London and J. Rosner, Phys. Rev. D34, 1530 (1986).

[88] W. J. Marciano, “Spin and Precision Electroweak Physics,” Proceedings of the 21st

SLAC Summer Institute, (1993).

[89] P. Langacker et al., Phys. Rev. D30, 1470 (1984).

[90] F. Abe et al. (CDF Collaboration), Phys. Rev. Lett 79, 2192 (1997).

[91] M. E. Peskin and T. Takeuchi, Phys. Rev. D46, 381 (1991).

[92] I. Maksymyk et al., Phys. Rev. D50, 529 (1993).

[93] A. Kundu and P. Roy, Int. J. Mod. Phys. A12, 1511 (1997).

[94] C. P. Burgess et al., Phys. Lett. B326, 276 (1994).

[95] D. S. Armstrong et al., hep-ex/0308049 (2003).

Page 193: A Precision Measurement of the Weak Mixing Angle in Møller ... · APV = −158 ppb±21 ppb (stat) ±17 ppb (sys). (1) The result represents the first demonstration of parity violation

170

[96] J. Arrington et al., “Parity Violating Deep Inelastic Scattering at CEBAF 6 GeV,

Letter of Intent to Jefferson Lab PAC-24,” Jefferson Lab internal document (2003).

[97] J. Arrington et al., “Hall A 12 GeV Upgrade Pre-Conceptual Design Report,” Jefferson

Lab internal document (2003).

[98] G. Azuelos et al., hep-ph/0003275 (1999).

[99] A. Czarnecki and W. J. Marciano, Int. J. Mod. Phys. A15, 2365 (2000).

[100] S. Kuhlman et al., “Physics and Technology of the Next Linear Collider,” hep-

ex9605011 (1996).